148
HAL Id: tel-01930812 https://tel.archives-ouvertes.fr/tel-01930812 Submitted on 22 Nov 2018 HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers. L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés. Investigating the role of voltage-gated ion channels in pulsed electric field effects in excitable and non-excitable cell lines Ryan Burke To cite this version: Ryan Burke. Investigating the role of voltage-gated ion channels in pulsed electric field effects in excitable and non-excitable cell lines. Human health and pathology. Université de Limoges, 2017. English. NNT : 2017LIMO0118. tel-01930812

Investigating the role of voltage-gated ion channels in

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Investigating the role of voltage-gated ion channels in

HAL Id: tel-01930812https://tel.archives-ouvertes.fr/tel-01930812

Submitted on 22 Nov 2018

HAL is a multi-disciplinary open accessarchive for the deposit and dissemination of sci-entific research documents, whether they are pub-lished or not. The documents may come fromteaching and research institutions in France orabroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, estdestinée au dépôt et à la diffusion de documentsscientifiques de niveau recherche, publiés ou non,émanant des établissements d’enseignement et derecherche français ou étrangers, des laboratoirespublics ou privés.

Investigating the role of voltage-gated ion channels inpulsed electric field effects in excitable and non-excitable

cell linesRyan Burke

To cite this version:Ryan Burke. Investigating the role of voltage-gated ion channels in pulsed electric field effects inexcitable and non-excitable cell lines. Human health and pathology. Université de Limoges, 2017.English. �NNT : 2017LIMO0118�. �tel-01930812�

Page 2: Investigating the role of voltage-gated ion channels in

1

Université de Limoges École Doctorale Bio-Santé (ED 524)

Faculté de Sciences et Techniques – XLIM Équipe BioEM

Thèse pour obtenir le grade de

Docteur de l’Université de Limoges Discipline – Biologie, médicine et santé

Présentée et soutenue par

Ryan Burke

Le 19 décembre 2017

Thèse dirigée par Dr. Philippe Leveque

Co-dirigée par Dr. Sylvia Bardet

JURY :

Président du jury Mme. Catherine YARDIN Professeur des universités, Praticien hospitalier

Rapporteurs M. Justin TEISSIE Directeur de recherche CNRS, émérite M. Yann PERCHERANCIER Chargé de recherche CNRS, HDR

Examinateurs M. Mounir TAREK Directeur de recherche CNRS Mme Sylvia BARDET Maître de conférences, Université de Limoges M. Philippe LEVEQUE Directeur de recherche CNRS

Investigating the role of voltage-gated ion channels in pulsed electric field

effects in excitable and non-excitable cell lines

Thèse de doctorat

Page 3: Investigating the role of voltage-gated ion channels in

2

Structure of the thesis

All experiments described in this thesis were conducted over the period of 2014 – 2017.

They were carried out primarily in the Bioelectromagnetics lab at XLIM Research Institute in

Limoges, France. Another portion of the data presented in this thesis was the result of a

collaboration with the group at the Biocybernetics lab at the University of Ljubljana.

The thesis is outlined as follows:

Chapter 1 provides a review of the literature and describes the general methodology for

this thesis.

Chapter 2 is the first experimental section that employs advanced biostatistics to

examine how well data can be predicted using common parameters present in current

electrophysical models of electropermeabilization.

Chapter 3 looks at the effects of nanosecond pulsed electric fields on the

transmembrane potential in U87 glioblastoma cells. Using a host of pharmacological

modulators of ion channel activity, the role of voltage-gated ion channels is explored.

Chapter 4 is the result of a collaborative exchange with the University of Ljubljana. We

look at the effect of pulsed electric fields, ranging from 10 ns to 10 ms, on membrane

potential and membrane permeability of several cell lines.

Chapter 5 begins with a brief summary of each chapter, and then moves into a more

global discussion of how our results contribute to the current literature.

Page 4: Investigating the role of voltage-gated ion channels in

3

Acknowledgements

There are so many people that I wish to thank for their inspiration, support and guidance.

I would like to thank Dr. Michael Persinger who was my thesis supervisor during my

B.Sc. and M.Sc. candidacy. You have inspired me in ways you will never know. You

demanded excellence and creativity and pushed me to grow beyond what I imagined

possible. I am grateful, maybe above all, for your emphasis on developing a very strong

knowledge of biostatistics. There is no tool more important or more powerful in science than

the ability to properly measure and interpret experimental data.

Thank you to Dr. Philippe Leveque. Despite being overwhelmed with a tremendous

work load, you took me on as your student half way through my Ph.D. Thank you for all of

your guidance and support. If I could go back in time, I would have selected you as my

supervisor for my entire Ph.D.

I am grateful to several colleagues who I am happy to call friends. Lynn, Maarten and

Stine, you all have helped so much. What an honor to have worked with you. I wish you all

prosperity and happiness in your future endeavors.

Thank you to my amazing family. Mom, you are one of the most inspiring people I

have ever known. I am here because of you. Thank you for everything you have ever done for

your family. I cherish you. Dad, I can’t begin to express how much I miss you. You taught me

what it means to be a man. You taught me to pursue my dreams and to never give up. This

thesis is dedicated to your memory. Cancer took you from us, and I hope that my humble

contribution to the field of cancer research would have made you proud. Tom, you are such

an amazing brother. Although you’re my younger brother, I have always looked up to your

incredible work ethic. I am so proud of the man you have grown into. You deserve nothing

but excellence. Les, I hope you realize that I love you so much. You too are an amazing man

and I am so happy for you and your beautiful family. I have the best brothers ever! I love you

all!

Page 5: Investigating the role of voltage-gated ion channels in

4

To the love of my life, Melanie. You have no idea how much you mean to me. You put

all of your dreams on hold for me to be here today. When faced with the idea of selling

everything you own and flying across the ocean far away from your family friends, you didn’t

hesitate for one second. There is no luckier man than I. You are truly a remarkable woman! I

feel blessed for every moment we share and I look forward to the next chapter of our lives.

From the bottom of my heart, thank you.

Page 6: Investigating the role of voltage-gated ion channels in

5

Abstract (English)

The use of pulsed electric fields (PEF) in medical and biotechnology sectors has

become increasingly prevalent over the last few decades. Research has shown that by adjusting

the duration of the PEF we can predict what effects will be observed. Whereas PEF in the

micro-to-millisecond range have been used to permeabilize the cell membrane and enhance

drug or protein uptake, nanosecond PEF (nsPEF) have demonstrated unique effects on

intracellular organelles. Both PEF and nsPEF have demonstrated therapeutic potential for a

variety of human pathologies, including the treatment of cancer. Using live-cell imaging, this

thesis investigated, in vitro, the effects of pulsed fields ranging in duration from 10 ns to 10

ms on cancerous (U87 glioblastoma multiforme) and non-cancerous cell lines (mouse

hippocampal neurons (HT22) and Chinese hamster ovary (CHO) cells). Previously published

results have demonstrated that cancerous cells have a greater sensitivity to applied electric

fields than healthy cells do. Our results are in agreement with these findings, insofar as the

U87 cells underwent a significantly greater depolarization of their transmembrane potential

following a single electric pulse at all durations. In a parallel set of experiments, despite having

similar electric field thresholds for membrane permeabilization, the U87 cells demonstrated

significantly enhanced YO-PRO uptake compared to the other cells lines. Although U87 cells

underwent the greatest change in both membrane depolarization and membrane

permeabilization, they also showed the fastest membrane resealing constant, which was

approximately 30 seconds faster than other cell lines. To elucidate some of the underlying

mechanisms by which U87 cells respond to electric fields, a series of experiments looked at the

role of transmembrane ion channels. Several recent studies have reported that PEFs can act

directly on voltage-gated ion channels. Using a variety of specific and broad acting

pharmacological ion channel modulators, we demonstrated that we could almost entirely

inhibit the electric field-induced membrane depolarization in U87 cells by blocking certain

cationic channels. These results were quite specific, such that the big conductance potassium

(BK) channel, L- and T-type calcium channels, and the non-specific cationic channel, TRPM8,

were able to inhibit depolarization while blocking other ion channels produced no significant

Page 7: Investigating the role of voltage-gated ion channels in

6

change. The work in this thesis showed that the malignant U87 cell line showed a greater

sensitivity to electric fields from ranging from 10 ns – 10 ms when compared to the non-

cancerous cell lines that were investigated. Potential improvements to current treatment

protocols have been proposed based on the findings presented herein.

Page 8: Investigating the role of voltage-gated ion channels in

7

Résumé (Français)

L'utilisation de champs électriques pulsés (PEF) dans les secteurs de la médecine et de

la biotechnologie est devenue de plus en plus courante au cours des dernières décennies. La

recherche a montré qu'en ajustant la durée du PEF, nous pouvons prédire quels effets seront

observés. Alors que les PEF dans la gamme micro - milliseconde ont été utilisés pour

perméabiliser la membrane cellulaire et améliorer l'absorption de médicament ou de protéine,

le PEF nanoseconde (nsPEF) a démontré des effets uniques sur les organites intracellulaires.

Les deux PEF et nsPEF ont démontré un potentiel thérapeutique pour une variété de

pathologies humaines, y compris le traitement du cancer. Utilisant l'imagerie des cellules

vivantes, cette thèse a étudié in vitro les effets de champs pulsés d'une durée de 10 ns à 10 ms

sur des lignées cancéreuses (U87 glioblastome multiforme) et non cancéreuses (neurones

hippocampes de souris (HT22) et cellules ovariennes du hamster chinois (CHO)). Des

résultats publiés antérieurement ont démontré que les cellules cancéreuses sont plus sensibles

aux champs électriques que les cellules saines. Nos résultats sont en accord avec ces résultats,

dans la mesure où les cellules U87 ont subi une dépolarisation significativement plus

importante de leur potentiel transmembranaire après une seule impulsion électrique à toutes

les durées. Dans un ensemble d'expériences parallèles, malgré des seuils de champ électrique

similaires pour la perméabilisation membranaire, les cellules U87 ont démontré une

absorption significativement améliorée de YO-PRO par rapport aux autres lignées cellulaires.

Bien que les cellules U87 aient subi le plus grand changement dans la dépolarisation

membranaire et la perméabilisation membranaire, elles ont également montré la constante de

rescellement de la membrane la plus rapide, qui était environ 30 secondes plus rapide que les

autres lignées cellulaires. Pour élucider certains des mécanismes sous-jacents par lesquels les

cellules U87 répondent aux champs électriques, une série d'expériences a examiné le rôle des

canaux ioniques transmembranaires. Plusieurs études récentes ont rapporté que les PEF

peuvent agir directement sur les canaux ioniques voltage-dépendants. En utilisant divers

modulateurs de canaux ioniques pharmacologiques spécifiques et à action large, nous avons

Page 9: Investigating the role of voltage-gated ion channels in

8

démontré que nous pouvions presque entièrement inhiber la dépolarisation membranaire

induite par le champ électrique dans les cellules U87 en bloquant certains canaux cationiques.

Ces résultats étaient assez spécifiques, tels que le canal de potassium de grande conductance

(BK), les canaux calciques de type L et T, et le canal cationique non spécifique, TRPM8, étaient

capables d'inhiber la dépolarisation tandis que le blocage d'autres canaux ioniques ne

produisait aucun changement significatif. . Les travaux de cette thèse ont montré que la lignée

cellulaire maligne U87 présentait une plus grande sensibilité aux champs électriques allant de

10 ns à 10 ms par rapport aux lignées cellulaires non cancéreuses étudiées. Des améliorations

potentielles aux protocoles de traitement actuels ont été proposées sur la base des résultats

présentés ici.

Page 10: Investigating the role of voltage-gated ion channels in

9

Table of Contents

Structure of the thesis .......................................................................................................................... 2

Acknowledgements .............................................................................................................................. 3

Abstract (English) ................................................................................................................................ 5

Résumé (Français) ............................................................................................................................... 7

Table of figures ................................................................................................................................... 12

Chapter 1: General Introduction ................................................................................................ 15

1.1 Pulsed electric fields ..................................................................................................................... 16

1.2 Biological membranes ................................................................................................................. 16

1.3 Transmembrane proteins ............................................................................................................ 17

1.4 Describing the cell as an electrical circuit ................................................................................. 19

1.5 Electric field interaction with the cell membrane .................................................................... 20

1.6 Factors influencing outcome following membrane exposure to an electric field ................ 21

1.6 The role of pulse duration in PEF effects .................................................................................. 23

1.7 PEF effects and applications ....................................................................................................... 24

1.7.1 Irreversible electropermeabilization (IRE) ........................................................................ 24

1.7.2 Reversible electropermeabilization .................................................................................... 24

1.8 nsPEF effects and applications .................................................................................................. 25

1.9 Limitations of PEF/nsPEF as a clinical treatment .................................................................. 28

1.10 General methods ........................................................................................................................ 28

1.11 Culture and maintenance of cell lines ...................................................................................... 29

1.11.1 Cell lines and maintenance ................................................................................................. 29

1.11.2 Differentiation of HT22 cells.............................................................................................. 30

1.12 Exposure of cells to electric pulses ........................................................................................... 30

1.13 Fluorescence imaging ................................................................................................................ 32

1.14 Fluorescent indicators for measuring transmembrane potential and membrane

permeability ........................................................................................................................................ 33

1.14.1 Measuring membrane permeability .................................................................................. 33

1.14.2 Measuring transmembrane potential ............................................................................... 34

1.15 Image acquisition ....................................................................................................................... 38

1.16 Statistical analysis ...................................................................................................................... 39

1.16.1 Independence of data .......................................................................................................... 39

1.16.2 Normal distribution ............................................................................................................ 40

1.16.3 Homogeneity of variance ................................................................................................... 40

1.16.4 Violations of statistical assumptions ................................................................................ 40

Page 11: Investigating the role of voltage-gated ion channels in

10

1.16.5 Additional inclusionary criteria......................................................................................... 41

1.16.6 Definition of threshold ....................................................................................................... 41

1.16.7 Statistical terminology and abbreviations ....................................................................... 41

1.17 Objectives and hypotheses for this Thesis ............................................................................... 42

Chapter 2: A comparative analysis of the theoretical and experimental interactions of PEF

with cells in vitro ................................................................................................................................ 44

2.1 Introduction .................................................................................................................................. 45

2.2 Methods ........................................................................................................................................ 48

2.2.1 Determining size, shape and orientation of cells to PEF ..................................................... 48

2.2.1.1 Calibration of images ......................................................................................................... 48

2.2.1.2 Determining size of the cell .............................................................................................. 49

2.2.1.3 Determining shape of a cell .............................................................................................. 49

2.2.1.4 Determining the angle of the cell with respect to the applied electric field .............. 49

2.2.1.5 Determining the density of cells in a given image ......................................................... 49

2.2.2 Statistical analyses ................................................................................................................... 50

2.3 Results ........................................................................................................................................... 50

2.4 Discussion ..................................................................................................................................... 54

Chapter 3: Nanosecond pulsed electric fields depolarize transmembrane poten-tial via

voltage-gated K+, Ca2+ and TRPM8 channels in U87 glioblastoma cells. (Based on published

manuscript) 57

3.1 Introductory Remarks ................................................................................................................. 58

3.2 Introduction ................................................................................................................................. 58

3.3 Materials and Methods ............................................................................................................... 61

3.3.1 Pharmacological manipulation of ion channel activity ................................................... 61

3.3.2 Calibration of PMPI voltage-dye ........................................................................................ 63

3.3.3 Statistical analyses................................................................................................................ 64

3.4 Results ........................................................................................................................................... 64

3.4.1 Calibration of PMPI .................................................................................................................. 64

3.4.1.1 Potassium calibration ........................................................................................................ 65

3.4.1.2 Calibration using electrophysiology ................................................................................ 66

3.4.2 Determining the threshold of electric field intensity required for nsPEF-induced

membrane depolarization. ................................................................................................................ 67

3.4.3 Fluorescence imaging of plasma membrane depolarization following a single 34 kV/cm

nsPEF. .................................................................................................................................................. 69

3.4.4 Inhibition of nsPEF depolarizing effect with BK channel blockers. .................................. 70

3.4.5 Depolarizing response of nsPEF is calcium-sensitive. ........................................................ 75

Page 12: Investigating the role of voltage-gated ion channels in

11

3.4.6 nsPEF-induced membrane depolarization is not mediated by voltage-gated Na+

channels; however, Na+ ions may still be involved. ....................................................................... 81

3.4.7 nsPEF-block is reversible. ....................................................................................................... 83

3.5 Discussion ..................................................................................................................................... 84

3.6. Conclusion ................................................................................................................................... 88

Chapter 4 - Plasma membrane depolarization and permeabilization due to electric

pulses in cell lines of different excitability ...................................................................................... 90

4.1 – Introductory remarks ............................................................................................................... 91

4.2 Introduction ................................................................................................................................. 92

4.3 Materials and Methods ............................................................................................................... 94

4.3.1 Cell culture and preparation ................................................................................................ 94

4.3.2 Potassium calibration of PMPI. .......................................................................................... 94

4.4 Results ........................................................................................................................................... 95

4.4.1 Cell Excitability ..................................................................................................................... 95

4.4.2 Plasma Membrane Permeability ........................................................................................ 99

4.5 Discussion ................................................................................................................................... 101

4.6 Conclusion .................................................................................................................................. 104

Supplementary Data ........................................................................................................................ 106

Chapter 5 – Discussion and Conclusion ...................................................................................... 108

5.1 Summary ..................................................................................................................................... 109

5.2 A comparative analysis of the theoretical and experimental interactions of PEF with cells

in vitro ............................................................................................................................................... 109

5.3 Nanosecond pulsed electric fields depolarize transmembrane potential via voltage-gated

K+, Ca2+ and TRPM8 channels in U87 glioblastoma cells .......................................................... 110

5.4 Plasma membrane depolarization and permeabilization due to electric pulses in cell lines

of different excitability .................................................................................................................... 112

5.5 Conclusion .................................................................................................................................. 113

5.6 Perspectives ................................................................................................................................ 115

References ......................................................................................................................................... 117

List of publications during Ph.D. candidature ............................................................................. 145

Journal articles ............................................................................................................................. 145

Presentations at International Conferences ............................................................................. 146

Page 13: Investigating the role of voltage-gated ion channels in

12

Table of figures

Figure 1.2 – Ion channels are classified by the stimulus which modulates their activity ........ 18

Figure 1.3 – The equivalent electric circuit model of the cell membrane. ................................. 19

Eq 1.1 Schwan equation (steady-state). ........................................................................................... 20

Figure 1.4 – The induced membrane potential varies along the cell membrane. ..................... 21

Figure 1.5 – Results from a case study of a patient with malignant melanoma. ....................... 26

Figure 1.6 – In-human trial for treatment of basal cell carcinoma with nsPEF. ....................... 27

Figures 1.7 (a-b) Photos of the electrodes, used for experiments presented in this thesis. ..... 31

Figure 1.8 – Fluorescence imaging setup. ...................................................................................... 32

Figure 1.9 – Direct comparison of YP and PI uptake following nsPEF exposure. .................... 33

Figure 1.10 – Fast-response vs slow-response voltage-probes. ................................................... 34

Table 1.1 – Describing fast vs slow voltage probes. ....................................................................... 35

Figure 2.1 – Schwan equation (steady - state). .............................................................................. 45

Equation 2.1 - Schwan equation (first - order). ............................................................................. 46

Figure 2.2 – Calcium wave entering the cell from the anodic pole following nsPEF. .............. 47

Figure 2.3 – Theoretical interaction of an electric field and a cell membrane. ......................... 48

Equation 2.2 - Determining the shape of a cell. ............................................................................. 49

Equation 2.3 – Equation describing relationship between cell shape and density with PEF-

induced membrane depolarization. ................................................................................................. 51

Figure 2.4 - Investigating pattern in baseline fluorescence. ........................................................ 52

Table 2.1 - Descriptive statistics for the two populations of cells grouped by baseline

fluorescence. ....................................................................................................................................... 53

Table 2.2 - Pearson correlation table from the multiple regression test divided into clusters.

.............................................................................................................................................................. 54

Table 3.1 - List of pharmacological agents used throughout this investigation. ....................... 63

Figure 3.1 - Calibration of PMPI by varying extracellular K+. ..................................................... 65

Page 14: Investigating the role of voltage-gated ion channels in

13

Equation 3.1 – Equation describing relationship between extracellular potassium

concentration and membrane depolarization. ............................................................................... 66

Equation 3.2 – Equation describing relationship between membrane potential and PMPI

fluorescence. ....................................................................................................................................... 66

Equation 3.3 – Goldmann equation. ............................................................................................... 67

Table 3.2 - Comparison between calibration methods for PMPI. ............................................... 67

Figure 3.2- Determining the electric field threshold for a single nsPEF to depolarize plasma

membrane. .......................................................................................................................................... 68

Figure 3.3 - Fluorescence imaging of U87 cells prior to and after delivering a single nsPEF

with an electric field intensity of 34 kV/cm. ................................................................................... 69

Figure 3.4- Effects of potassium channel blockers on nsPEF-induced membrane

depolarization. .................................................................................................................................... 71

Figure 3.5 - Concentration-dependent inhibition of nsPEF depolarization by TEA. ............... 72

Figure 3.6 - BK channel blockers significantly inhibit membrane depolarization following 34

kV/cm nsPEF. ..................................................................................................................................... 74

Figure 3.7 - Calcium-dependent threshold response curves in response to nsPEF of varying

electric field strengths. ...................................................................................................................... 76

Figure 3.8 - Comparing the role of calcium from intracellular, extracellular and endoplasmic

reticulum compartments. .................................................................................................................. 78

Figure 3.9 - Effects of calcium channel blockers on nsPEF-induced membrane

depolarization. .................................................................................................................................... 80

Figure 3.10 - Comparing effects from sodium and chloride channel blockers on nsPEF-

induced membrane depolarization. ................................................................................................. 82

Figure 3.11 - Reversibility of the TEA and Penitrem A inhibition of the nsPEF-induced

membrane depolarization. ................................................................................................................ 83

Figure 3.12 - Proposed mechanism of direct interaction between nsPEF and voltage-gated

channels along with downstream effects on non-voltage dependent channels. ....................... 86

Page 15: Investigating the role of voltage-gated ion channels in

14

Figure 4.1 - Phase-contrast images of all four cell lines used in experiments. .......................... 95

Figure 4.2 – Chemical depolarization of cells using K+. .............................................................. 96

Figure 4.3 – Representative depolarization dynamics following PEF exposure. ...................... 97

Figure 4.4 – Magnitude of depolarizing response to pulsed fields from 10 ns – 10 ms. .......... 97

Table 4.1 - The depolarization thresholds for all tested pulse durations and cell lines............ 98

Figure 4.5 - The strength-duration curve for depolarization thresholds of all cell lines. ........ 99

Figure 4.6 - Normalized permeabilization curve of all four cell lines to YO-PRO, 5 min after

the pulse application. ....................................................................................................................... 100

Figure 4.7 - Time dynamics of YO-PRO uptake and analyses. .................................................. 101

Table 4.2 - Parameters of the fitted symmetric sigmoid to the normalized data of YO-PRO

uptake. ............................................................................................................................................... 101

Table A1 - Statistical parameters for the strength-duration curve by cell line. ....................... 106

Table A2 - Additional statistical parameters from the strength-duration curve. .................... 107

Page 16: Investigating the role of voltage-gated ion channels in

15

Chapter 1:

General Introduction

Page 17: Investigating the role of voltage-gated ion channels in

16

1.1 Pulsed electric fields

The last few decades have seen an increased interest in the study of pulsed electric

fields, which has been particularly evident in medical and biotechnology sectors. They have

been shown to have broad applications in the food processing industry [1–3], for

cryopreservation of cell lines [4–6], enhancing gene or drug uptake by cells [7–10], and induce

cell death through necrotic or apoptotic pathways [11–13].

Understanding how electric fields influence cell behavior is important when

considering how to improve or expand current treatment options. The following section will

begin by taking a closer look at the cell and its intracellular contents, but more specifically at

the membranes that surround them, where electric fields are expected to interact.

1.2 Biological membranes

Understanding how a cell interacts with its environment begins with the very boundary

that separates them, which of course is its membrane. The cell membrane (Fig. 1.1) is

composed of a series of polar lipid molecules that have hydrophobic and hydrophilic regions

organized into a bilayer [14]. This structure displays self-organizing properties in an aqueous

environment that results in a boundary condition separating the cytoplasm from extracellular

fluids. A very similar arrangement of amphipathic lipid molecules within the cell serves to

separate organelles from the cytoplasm. These membranes are critical for selective transport

into and out of the cell.

In addition to the phospholipid bilayer, a host of other lipids, proteins and

carbohydrates can be identified. These molecules are highly variable in their distribution along

the membrane, each serving different functions ranging from signaling molecules to structural

support to name a couple. Transmembrane protein channels are of particular interest to our

topic and will be discussed in more detail.

Page 18: Investigating the role of voltage-gated ion channels in

17

Figure 1.1 – Structure of cell membrane. The cell membrane is a highly complex

structure whose fundamental unit is the phospholipid molecule (bottom right of the figure).

The amphipathic structure of this molecule results in the formation of a bilayer in aqueous

solution. Embedded within the membrane, multiple additional lipids, proteins and

carbohydrates serve to communicate with the external environment. Figure taken from:

https://www.thinglink.com/scene/634047922140348418.

1.3 Transmembrane proteins

Transmembrane proteins, to a large degree, dictate the function of the cell. Some serve

as signal conduits, whereby extracellular molecules interact with the protein leading to a

specific intracellular effect [15]. Others serve as selectivity filters responsible for modulating,

either passively or actively, the passage of ions. These ion channels are highly variable and are

critical for many cell processes.

Ion channels fluctuate between open and closed states, and these states can be

influenced through multiple mechanisms (Fig 1.2). Some examples include voltage-gated

channels, mechanically-gated channels, or ligand-gated channels. Upon activation, these

channels become permeable to specific ions. They can be highly specific, meaning they allow

just one type of ion to pass, or charge-specific allowing either cations or anions to pass.

Page 19: Investigating the role of voltage-gated ion channels in

18

Figure 1.2 – Ion channels are classified by the stimulus which modulates their

activity. Voltage-gated channels can be activated by fluctuations in a cells transmembrane

potential. Ligand-gated channels are activated through signal molecules that bind on extra- or

intracellular sites of the channel. Mechanically-gated channels are modulated through

changes in shape associated a physical stimulus. Image taken from Molecular biology of the

cell, 4th edition at https://www.ncbi.nlm.nih.gov/books/NBK26910/.

The selective nature of these ion channels is responsible for the generation of a

transmembrane potential, which is equally present and varies among intracellular organelles

as it is across the plasma membrane. The chemical basis of this voltage-gradient, known as the

resting membrane potential, is the uneven distribution of ions inside and outside of the cell

[16]. The electrical properties of ions, combined with the conductive properties of protein

channels and the insulating properties of the membranes lipid bilayer, have allowed the cell

to be modeled as an electric circuit. This has been useful when studying electrical

communication between cells for example, but also when developing treatment strategies to

modulate cell activity.

Page 20: Investigating the role of voltage-gated ion channels in

19

1.4 Describing the cell as an electrical circuit

In the 1950’s Alan Hodgkin and Andrew Huxley proposed a model to describe the ionic

mechanisms responsible for the propagation of electrical impulses along a giant squid axon.

As a general overview (Fig. 1.3), the Hodgkin-Huxley model represents the lipid bilayer as

parallel capacitors capable of storing charge and ion channels as variable resistors capable of

passing current. The transmembrane potential is generated across the cell membrane and

fluctuates according to the activity of the ion channels.

This model has been invaluable in understanding certain biological phenomena such

as electrochemical communication between cells, particularly between electrically excitable

cells in nervous and muscular tissues. Another application for this model extends into fields

interested in modulating cellular activity using applied electric fields.

Figure 1.3 – The equivalent electric circuit model of the cell membrane. In this

model, the lipid bilayer is represented as a pair of parallel capacitors and the transmembrane

ion channels are represented as variable resistors. Image adapted from Lehigh University

Bioengineers page, https://lehighbioe.wordpress.com/2012/10/23/the-cell-membrane-as-a-

circuit/.

Page 21: Investigating the role of voltage-gated ion channels in

20

1.5 Electric field interaction with the cell membrane

When cells are exposed to an electric field, an electric force is generated, which acts on

ions both inside the cell and in the external media. Because ions are charged molecules, this

force causes them to move along the electric field lines. Whereas the extra- and intracellular

solutions are conductive, the cells lipid membrane is non-conductive. As a result these ions

will accumulate along the membrane and generate a large transmembrane potential. At a

certain point, which is approximately 1 V [17, 18], the induced voltage exceeds the membrane

capacitance and breakdown of the membrane occurs. When considering the diameter of the

membrane is approximately 10 nm, the electric field strength associated with this threshold is

on the order of a MV/cm. This process, which is termed electroporation or

electropermeabilization is associated with enhanced membrane permeability.

To describe this effect, the Schwan equation (Eq. 1.1) was developed. This equation is

limited to modeling a spherical cell, and states that the induced membrane potential (Vm) is

proportional to the radius of the cell (R) and the applied electric field (E), and will not be

uniform along the cell membrane (Fig. 1.4). Extensions of this model have been developed to

help understand more complex models, such as irregular-shaped cells that can’t be accurately

described as spherical or ellipsoid.

𝑉𝑚 =3

2 𝐸𝑅 𝑐𝑜𝑠𝜃

Equation 1.1 Schwan equation (steady-state). This describes the effect of an applied

electric field (E) on the transmembrane potential (Vm) given a cell radius of (R) as a function

of the angle (θ).

Page 22: Investigating the role of voltage-gated ion channels in

21

Figure 1.4 – The induced membrane potential varies along the cell membrane.

This is an example of the Schwan equation applied to a spherical cell with a radius of 10 µm

and an applied field intensity of 44 kV/cm.

More recently, there has been a shift in methodology toward understanding the

biomolecular events occurring when an electric field interacts with a lipid bilayer. Molecular

dynamics (MD) simulations have helped to clarify some of the events associated with

electropermeabilization, and have allowed us to visualize the dynamics associated with

membrane breakdown [19–23]. Up to now, limitations in the computational power have

restricted MD to a period less than 1 millisecond following pulse delivery [24]; however, a

multitude of effects have been reported over much longer time periods. These effects appear

to be highly dependent on several factors such as; pulse duration, electric field strength,

number of pulses, and frequency of pulses delivered.

1.6 Factors influencing outcome following membrane exposure to an electric field

It is generally accepted that the initial step in membrane breakdown following PEF

exposure is the magnitude of the induced transmembrane potential [17, 18]. Thus, it is

unsurprising that electric field strength would play an important role in this process. Finding

-0.8

-0.6

-0.4

-0.2

0

0.2

0.4

0.6

0.8

0 50 100 150 200 250 300 350 400

Ind

uce

d V

m (

V)

Degrees

Page 23: Investigating the role of voltage-gated ion channels in

22

the threshold for this response has been the subject of many investigations seeking to exploit

this consequence of PEF application. Achieving this threshold; however, is highly dependent

on the duration of the pulse delivered, such that eliciting a response using a shorter pulse

would require a stronger electric field than the same response using a longer pulse. For

example, using pulses ranging from 1 ns to 100 ms to stimulate gastrocnemius muscles, an

applied voltage of 4.5 kV was required for a 1.8 ns pulse to elicit a similar response to a 100

ms pulse delivered at 0.02 kV [25].

In addition to the electric field intensity and pulse duration, other important features

are; the number of pulses and the frequency at which they are delivered. These factors were

explored in [26], where nsPEF were investigated for their ability to cause swelling in Jurkat

cells. Single nsPEF were compared to 10 & 50 nsPEF. After 20 seconds significantly

enhanced swelling was seen as a function of number of pulses. In the same study, frequency

of pulses was explored, where swelling was measured after exposure to 10 pulses at 1 Hz or

10 pulses at 1 kHz. Similarly, significantly more swelling was observed in the group of cells

exposed to 10 nsPEF at 1 kHz when compared to the 1 Hz and the control groups.

In a study looking at Yo-Pro uptake in U87 cells following nsPEFs [27] the authors

report a significant enhancement in Yo-Pro uptake as a function of both number & frequency

of pulses delivered. Here, a single 10 nsPEF (1 p) was compared to 10 pulses (10 p) at 1 Hz,

10 Hz, and 100 Hz, and 100 pulses (100 p) at 1 Hz, 10 Hz, and 100 Hz. The results can be

summarized as follows: The greatest Yo-Pro uptake was observed – 100 p 100 Hz > 100 p 10

Hz > 100 p 1 Hz > 10 p 100 Hz > 10 p 10 Hz > 10 p 1 Hz > 1 p > control.

A noteworthy feature is evident when considering pulse shape in the context of the

literature presented above. In [28], using µsPEFs it was reported that the electric field

intensity required to permeabilize 50 % of the cells was reduced by 20 % when the authors

used symmetrical bipolar pulses. The working hypotheses for these results propose that i)

the polar asymmetry following a unipolar pulse is counterbalanced when a bipolar pulse is

delivered and ii) bipolar pulses increase the odds that a non-spherical cell are permeabilized.

Page 24: Investigating the role of voltage-gated ion channels in

23

In contrast to this research, several articles show a dampened effect when bipolar

nanosecond pulses are used. In [29], the addition of a second phase (of opposite polarity)

cancelled the effect of the initial monopolar phase in both intracellular calcium mobilization

and cell survival. These results persisted in spite of a doubling of total energy delivered to

cells, and was not dependent on pulse duration (60 ns or 300 ns), amplitude (15 – 60

kV/cm) or number of pulses delivered (1 – 60). This dampening effect gradually decreased as

the delay between the bipolar pulses increased. A similar dampening effect was observed for

calcium uptake [30] as well as measures of plasma membrane integrity Calcium Green-1,

Propidium Iodide and FM1-43 [31] when comparing bipolar effects to monopolar effects.

Although some explanations were put forth to explain the differences between µs and

ns bipolar pulses, the duration of the pulse is critical and is based on a biological constant

known as the membrane charging constant, which will be discussed in further detail in the

next section.

1.6 The role of pulse duration in PEF effects

Recall that the induced transmembrane potential is due to the accumulation of ions

along the cells membrane. This process obviously requires time, and is referred to as the

charging time, or charging constant. Although it varies slightly between cells, the charging

time is on the order of a few hundred nanoseconds [32–34]. Pulse durations that are longer

than the charging constant will have very different effects those that are shorter.

The following sections will examine the different effects expected when using pulsed

electric fields of various durations. For simplicity, pulses that are longer than the charging

time will be referred to as pulsed electric fields (or PEF), and those that are shorter will be

referred to as nanosecond pulsed electric fields (or nsPEF).

Page 25: Investigating the role of voltage-gated ion channels in

24

1.7 PEF effects and applications

At longer durations, in the micro- or millisecond range, PEF have been shown to be

effective at delivering drugs, proteins, and DNA across the cell membrane [35–41]. By

adjusting the number, frequency and intensity of the pulses one can reversibly or irreversibly

permeabilize the plasma membrane.

1.7.1 Irreversible electropermeabilization (IRE)

Irreversible electropermeabilization (IRE) has been studied as a purely electrical

treatment for tumors [13, 42–45]. Here, the cell is unable to repair the damage imposed by the

electric field. A review of this technique [46] looked at results from 106 patients over a period

ranging from 3 – 18 months. Efficacy of treatment ranged from 67 – 100% over this period.

Most recently, another human trial has shown promising results [47]. In this study, 30

patients were treated with liver tumors. They were monitored for 6 months following

treatment where treatment success was defined as no evidence of residual tumors in the

ablated area from CT and ultrasound scans. After 3 and 6 months, success was achieved in

79% and 66% of the patients, respectively.

1.7.2 Reversible electropermeabilization

Contrary to IRE, reversible electropermeabilization utilizes the ability of the

membrane to repair itself as an advantage. Of the two, reversible electropermeabilization has

been most thoroughly investigated, and offers a greater number of potential therapies. Gene

electrotransfer (GET) is one such application, where DNA is introduced into the cell and makes

it possible to regulate the production of a desired transgene. A very thorough review [39]

highlights the vast number of disorders that can potentially be treated, including Parkinson’s

disease, HIV/AIDS and cancer to name a few.

A more recent review of clinical GET trials highly the broad potential for this therapy

[48]. In this review, the author summarizes the results of more than 50 clinical trials (some of

which are ongoing) where GET has proven to be a safe and effective method to treat various

Page 26: Investigating the role of voltage-gated ion channels in

25

cancers by stimulating immune activity or disrupting angiogenesis in tumors; as well as a role

in DNA vaccines for HIV, Hepatitis, Human papilloma virus, and Hantaviruses.

Electrochemotherapy (ECT) is another application for reversible

electropermeabilization, where chemotherapeutics (or high concentrations of calcium) can be

injected systemically or locally followed by electropulsation of the tumor. Here drugs such as

bleomycin and cisplatin have been selected for their combination of low membrane

permeability and high cytotoxicity, and when combined with electropulsation, the toxicity of

these agents has been significantly enhanced [49–51].

Recently, a case study looked at the effects of calcium electroporation and

electrochemotherapy on a patient with malignant melanoma [52]. This patient had numerous

metastases (Fig. 1.5) yet 12 months following treatment complete leveling was observed in

both treated and non-treated lesions. The fact that non-treated tumors responded highlights

the potential role of the immune system in mediating this response. Biopsies taken upon

completion of the follow-up period revealed an absence of malignant cells.

1.8 nsPEF effects and applications

When pulse durations shorter than the charging constant are applied, the plasma

membrane no longer shields the intracellular environment [32, 33], and very different effects

can be expected. Effects such as mitochondrial depolarization [53, 54], caspase activation [55,

56] and nuclear condensation [57, 58]; all effects typical of cell death by apoptosis, have been

reported. Additional effects such as cytoskeleton disruptions, including microtubule [27] and

actin [59] assembly have also been reported.

Page 27: Investigating the role of voltage-gated ion channels in

26

Figure 1.5 – Results from a case study of a patient with malignant melanoma. The

top two images were taken prior to treatment using electrochemotherapy and calcium

electroporation. In these images multiple cutaneous metastases are evident. The bottom two

images were taken 16 months following treatment. A complete leveling was observed in both

treated and non-treated metastases. Biopsies of the pigmented lesions were found to be non-

malignant. Results and images from [52]

Whereas intracellular effects are specific to nsPEF, there is a lot of evidence suggesting

additional effects occur on the plasma membrane. For example, membrane permeability was

found to be enhanced following nsPEF exposure, where YO-PRO [27] and bleomycin [60]

uptake into the cell was observed. In an experiment using giant unilamellar vesicles, a single

nsPEF resulted in the delivery of siRNA molecules into the cytoplasm [61].

Page 28: Investigating the role of voltage-gated ion channels in

27

It has been hypothesized that these ultra-short fields may exert effects directly on

transmembrane protein channels in the plasma membrane [62]. Experimental evidence exists

that supports this, where nanosecond pulses have resulted in direct activation of voltage-gated

channels [63–65].

While nsPEF have demonstrated very unique effects compared to PEF, there is still a

need for clinical data to validate the translation of these effects in human trials. The first

human trial using nsPEF for cancer treatment was published in 2014 [66]. In this study a total

of 10 tumors were treated with nsPEF on 3 patients with basal cell carcinoma. Of the tumors

treated, seven were completely free of cancerous cells, two had partially regressed, and one

recurred 10 weeks after treatment. An example of tumor treatment in this study can be seen

in Fig. 1.6 below.

Figure 1.6 – In-human trial for treatment of basal cell carcinoma with nsPEF. An

example from [66], where the first human trial was conducted testing nsPEF efficacy on skin

cancer. Three patients were treated using nsPEF and followed for several weeks following

treatment. Images above are a representative example of the effects observed. The first and

second images were taken pre- and immediately post-nsPEF exposure. The final image was

taken 10 weeks post-treatment where there was no evidence of tumor recurrence. This was

confirmed by histological analysis of biopsy.

Despite all of the promising research using PEF and nsPEF, this field is still continuing

to evolve. Many questions still remain with respect to the mechanism of action of PEF/nsPEF

Page 29: Investigating the role of voltage-gated ion channels in

28

effects on biological systems. Understanding these mechanisms will help us overcome some

of the limitations present in this field of study.

1.9 Limitations of PEF/nsPEF as a clinical treatment

The major limitations facing experimentation using pulsed electric fields (PEF/nsPEF)

for treating humans are time and money. An abundance of research demonstrates the

potential of electrically based therapies. Recently, the handbook of electroporation was

published [67] containing a comprehensive series of chapters from various contributing

authors highlighting studies from modeling and simulations to in vitro work right up to

clinical applications. Despite the amount of research, very few countries offer treatments using

pulsed electric fields.

As a result, patients who could benefit from these treatments aren’t even being given

the option. And those who may be aware will have to make the choice between paying from

their own pockets and choosing those treatment options that are covered by health insurance

policies. That is why I say time and money are the major limitations. More money is required

to generate further research, and this takes time. With that said, it is an honor contribute in

the most humble way to this ever-growing field hoping that I can add to the current body of

knowledge.

The pulse generators, as well as the fluorescent indicators of transmembrane potential

and membrane permeability, used for these experiments will be discussed in detail in the

following section.

1.10 General methods

The following section will provide a detailed overview of some of the techniques and

equipment used throughout the work presented in this thesis. Attention will be given to

explaining the cell lines used and how they were maintained; the different pulse generators

and electrodes used; the fluorescence microscopes used and which fluorescent indicators were

selected; the methods involved in extraction of data from images, and the statistical

Page 30: Investigating the role of voltage-gated ion channels in

29

methodology applied. Part of the work described in this thesis was the result of a collaboration

with the University of Ljubljana. My contribution to this collaboration involved data collection

and statistical analyses. Whereas we investigated the role of pulse durations ranging from 10

ns to 10 ms, the majority of the work presented in this thesis focuses on 10 nsPEF, which is

the primary focus of our lab.

1.11 Culture and maintenance of cell lines

1.11.1 Cell lines and maintenance

Three cell lines were used for the experiments that will be presented. Chinese hamster

ovary (CHO) cells (European Collection of Authenticated Cell Cultures ECACC, cells CHO-K1,

cat. no. 85051005, obtained directly from the repository), U87-MG human glioblastoma cells

(ECACC, Public Health England, cat. no. 89081402) and HT22 immortalized mouse

hippocampal neurons (The Salk Institute, La Jolla, CA) were grown in an incubator at 37 °C

at 5 % CO2 for 2-5 days before experiments. HT22 cells were used in their non-differentiated

and differentiated states. The procedure used to differentiate these cells will be explained

further in section 1.11.2

Experiments were completed in laboratories in Limoges, France and Ljubljana,

Slovenia. In Limoges, depending on the experimental setup, cells were grown in one of two

ways. They were either grown on Poly-Lysine (Sigma-Aldrich, Germany) coated 22 mm glass

coverslips in 35 mm Petri dishes (VWR International, USA); or they were grown in 35 mm

glass FluoroDishes (World precision instruments, USA). In the former condition, the 22 mm

glass coverslip was transferred into a customized plastic ring which could be mounted onto

the microscope. In Ljubljana, cells were grown in 40 mm Petri dishes (TPP, Austria). To ensure

these subtle differences in methodology were not significantly influencing the results,

experimental results were compared between labs.

For growth and maintenance of cell lines, CHO cells were grown in Ham-F12 media

(Sigma-Aldrich, Germany), U87 cells in MEM (Sigma-Aldrich, Germany) and the HT22 cells

Page 31: Investigating the role of voltage-gated ion channels in

30

in DMEM (Sigma-Aldrich, Germany) growth medium. All growth media were supplemented

with 10% fetal bovine serum (Gibco, France or Sigma-Aldrich, Germany, cat. no. F9665), L-

glutamine (StemCell, Canada) and antibiotics.

1.11.2 Differentiation of HT22 cells

HT22 cells were used in their non-differentiated and terminally differentiated forms.

In order for HT22 cells to be differentiated, DMEM growth media had to be replaced with

Neurobasal media (Thermofisher, USA, cat. no. 21103-049), and supplemented with 0.5 mM

L-glutamine and 1 mL B-27 (Thermofisher, USA, cat. no. 17504-044) for every 50 mL

Neurobasal media. Cells were maintained in this differentiation medium for 24 – 48 hours

prior to experimentation.

1.12 Exposure of cells to electric pulses

Cells were exposed to different pulse durations (10 ns, 550 ns, 1 µs, 10 µs, 100 µs, 1 ms,

and 10 ms). To accommodate experiments using these pulse durations, three different pulse

generators were used. For 10 ns pulses, a commercially available nsPEF generator (FPG 10-

1NM-T, FID Technology, Germany) with 50 Ω output impedance was used. The 10 ns pulse

durations (13.6 ns full width at half magnitude (FWHM) duration and a 5.2 ns rise time) were

applied with electric field intensities ranging from 16.5 – 52.0 kV/cm. A high-voltage

measurement device (tap-off) connected to a digital phosphor oscilloscope (DPO 4104,

Tektronix, USA) was used to visualize the time-domain measurements of each pulse. Pulses

were delivered to cells by positioning the electrodes (Fig. 1.7 a), comprised of two stainless-

steel electrodes (1.2 mm gap) with a 50 Ω impedance resistive load in parallel, using a

micromanipulator (MP28, Sutter Instruments USA). The applied electric field was

numerically determined [27]. Numerical modelling of the electrode delivery system was

performed with a Finite-Difference Time-Domain (FDTD) based electromagnetic solver, and

a finest spatial mesh of 100 μm.

Page 32: Investigating the role of voltage-gated ion channels in

31

For pulse exposures of 550 ns and 1 µs, a laboratory prototype nanosecond generator

with H-bridge MOSFET amplifier architecture and additional peripheral circuitry for voltage

adjustments and safety was used. Voltage could reach 1 kV with an accuracy of 3 %. For pulse

exposures of 10 µs – 10 ms a commercially available BetaTech electroporator (Electro cell B10,

Betatech, France) was used, and the current was measured with either current probe CP030

(Teledyne LeCroy, USA) or a Pearson current monitor (Model 2877, Pearson Electronics Inc.,

USA). Pulses longer than 10 ns were applied using either stainless-steel wire electrodes with a

4 mm intra-electrode distance and 1.29 mm diameter (16 gauge needle) or Pt/Ir wire

electrodes with 5 mm intra-electrode distance and 0.75 mm diameter (Fig. 1.7 b). The

numerical calculations show that the electric field in the middle of the electrodes is

approximately homogeneous [68]. To compare the results obtained with both electrodes, we

approximated the electric field between the electrodes as E = U / d, where U is the applied

voltage and d the distance between the electrodes. We adapted the applied voltage to achieve

the same electric field and acquired images in the middle of the two electrodes.

Figures 1.7 (a-b) Photos of the electrodes, used for experiments. A - Electrodes used

for delivering 10 ns pulses; B - electrodes used for delivering pulses longer than 500 ns.

Respective traces for 10-ns and 1-µs pulses delivered are to the right of electrode images, along

with a numerical simulation of the nsPEF field distribution.

Page 33: Investigating the role of voltage-gated ion channels in

32

1.13 Fluorescence imaging

Experiments conducted in Limoges were performed on a DMI6000 confocal

microscope (Leica Microsystems, Germany) using a 63x oil immersion objective (for

experiments measuring membrane potential) or a 20x objective (for experiments measuring

membrane permeability). A Spectra 7 light engine (Lumencor, USA) was used for fluorescence

excitation, whereas the emitted light was captured with an electron-multiplying charge-

coupled device camera (EMCCD Evolve 512, Rope, USA). Fig. 1.8 shows the imaging setup.

Figure 1.8 – Fluorescence imaging setup. An image of the workstation where the nsPEF

experiments were performed.

In Ljubljana, an inverted microscope AxioVert 200 (Zeiss, Germany) was used for

experiments with a 20x (when measuring membrane permeability) or 40x (when measuring

membrane potential) objective. Samples were excited using a high-speed polychromator

(Visitron systems GmbH, Germany), and the emitted fluorescence was acquired with a

VisiCam 1280 CCD camera (Visitron, Germany).

Page 34: Investigating the role of voltage-gated ion channels in

33

1.14 Fluorescent indicators for measuring transmembrane potential and membrane permeability

1.14.1 Measuring membrane permeability

Selecting a fluorescent dye for measuring membrane permeability was straightforward

because there are only a few that are commonly used in the literature. Generally, membrane

integrity is measured using either propidium iodide (PI) or YO-PRO (YP). There are multiple

studies using these dyes; however, when directly compared [69], YP was found to be a more

sensitive measure for membrane permeability than PI (Fig. 1.9). This difference was attributed

to the size of the molecule, insofar as YO-PRO has much less steric hindrance associated with

its planar configuration. For that reason it was used for the permeability experiments

presented in this thesis. In all cases, it was used at a concentration of 1 µM.

Figure 1.9 – Direct comparison of YP and PI uptake following nsPEF exposure.

These results demonstrate the difference in sensitivities between YP and PI for measurement

of membrane integrity. Image taken from [69].

Page 35: Investigating the role of voltage-gated ion channels in

34

1.14.2 Measuring transmembrane potential

Selecting a voltage probe, on the other hand, required more consideration. Typically

voltage-sensors are divided into two slow- or fast-probes. Depending on cell type (excitable vs

non-excitable) or on the experimental objectives, both voltage-sensors have advantages and

disadvantages. A summary of their differences is available below in Table 1.1.

The major difference between slow and fast probes is the mechanism by which they

permit the measurement of Vm. (Fig. 1.10). Slow probes, also called Nernstian probes, undergo

Vm - dependent changes in their transmembrane distribution. These probes typically are

considered when looking for long-term changes (seconds to minutes) in membrane potential

that is due to alterations in ionic permeability of the membrane.

Figure 1.10 – Fast-response vs slow-response voltage-probes. A In response to a

change in the local electric field, fast-response probes undergo a change in the electronic

configuration. This structural change leads to changes in the fluorescent properties of the

molecule. B Slow-response probes undergo a transmembrane redistribution that is dependent

on the membrane potential. Source for the figure -

https://www.thermofisher.com/fr/fr/home/references/molecular-probes-the-

handbook/probes-for-membrane-potential/introduction-to-potentiometric-probes.html

Page 36: Investigating the role of voltage-gated ion channels in

35

Whereas the optical response time for slow probes prevents the study of single action

potentials in excitable cells such as neurons and muscle cells, the magnitude of response is

much greater than fast-response probes, in some cases approaching 100 x more sensitive (~

1% change per mV). This allows for the measurement of subtler changes in Vm over time.

Fast-probes, or electrochromic probes, are typically selected when the interest is in

very rapid changes in membrane potential. These probes undergo a change in the electronic

structure of the molecule when the surrounding electric field is changed. This change in

structure results in a change in fluorescent properties. Because the change in fluorescence does

not depend on translocation across the membrane, the optical response time is extremely fast.

Contrary to the slow probes, the cost of the rapid response is a significant reduction in

sensitivity (~ 2 – 10% change per 100 mV).

Slow Fast

How they work Membrane potential-dependent change in their transmembrane distribution results in change in fluorescence

A change in surrounding electric field results in a change in the electronic structure of the probe which changes the fluorescent properties of the molecule

Magnitude of change ~ 1% change per mV ~ 2-10% change per 100 mV

Optical response Suitable for detection of average membrane potential changes (from respiratory activity, ion-channel permeability, drug binding etc.)

Suitable for detection of transient, rapid changes in potential (ms) in neurons and cardiac cells

Examples Carbocyanines and rhodamines (cations); and oxonols (anions)

AminoNaphthylEthenylPyridium (ANEP) dyes; Förster resonance energy transfer (FRET) sensors; FluoVolt*

Table 1.1 – Describing fast vs slow voltage probes. Comparing the pros and cons of

various types of probe based on their mechanism of action, magnitude of change and optical

response. * FluoVolt has magnitude of change closer to 25% per 100mV.

Page 37: Investigating the role of voltage-gated ion channels in

36

A commercially available slow probe, known as plasma membrane potential indicator

(PMPI); a commercially available fast-probe, FluoVolt; and a genetically encoded fast probe,

ArcLight were tested and compared to determine how they responded to a single nsPEF.

Experiments were performed to determine which would be more appropriate for our

experiments.

ArcLight is part of the rapidly evolving family of genetically encoded voltage sensors.

This next generation technology is made possible due to a fusion of a voltage sensor found in

the sea squirt, Ciona intestinalis, with a fluorescent protein (GFP) found in the jelly fish,

Aequorea victoria [70]. DNA transfection into the host cell and subsequent expression is made

possible using a baculovirus which is unable to replicate in mammalian cells [71–73]. First

described in 2012, ArcLight has been considered a revolutionary step forward in fast-response

probes and genetically encoded voltage sensors in particular because of its relatively large

magnitude of response, which has been reported to approach 35% per 100 mV change in

transmembrane potential [70].

Despite this large magnitude of response when compared to other fast voltage-probes,

when ArcLight-transfected U87 cells were exposed to a 34 kV/cm nsPEF, the relative change

was quite minimal. With such a minimal response taken with the relative inconvenience

associated with the time taken to transfect the cells, this probe was not considered ideal for

the purpose of our intended experiments.

The second fast-response voltage-sensing probe, FluoVolt, is considered to be the next

generation for electrochromes not included in the genetically encoded family. Because of the

commercial nature of this voltage-sensitive dye, the chemical structure is unknown.

Lifetechnologies, the company who distributes this dye, states that its optical resolution is in

the sub millisecond timescale and exhibits a magnitude of response up to 25% per 100 mV

change in membrane potential.

Although easy to use, requiring a simple mixing procedure along with a 15 – 30 minute

incubation period, the response observed following a 34 kV/cm nsPEF was similar to that

Page 38: Investigating the role of voltage-gated ion channels in

37

achieved with ArcLight. A recent article [65] using nanosecond pulsed electric fields confirmed

that the resultant response of FluoVolt, in this case to a 3 kV/cm, 200 ns PEF was closer to 5%

than the 25 % expected. This level of response would not be adequate to determine potentially

subtle effects following application of various ion channel modulators.

The commercially available slow-response probe, PMPI, was determined to be the

most capable for answering the questions we set out to answer, and will be addressed in the

following section. This probe, also of unknown chemical structure because of its commercial

nature, was able to reliably respond with a magnitude around 20 - 30x greater than the other

probes tested. Because of the kinetics of this dye, such that an increase in fluorescence is

associated with a depolarization of the transmembrane potential, the chemical family could be

determined.

Three types of slow-response probes exist: Carbocyanines, rhodamines, and oxonols.

The two former families are cationic species. Because they are positively charged, these

molecules tend to be used more for monitoring mitochondrial membrane potential whose

resting potential is significantly more negative than that of the cell’s plasma membrane [74,

75]. Oxonol dyes, on the other hand, are anionic species which do not label mitochondria or

other intracellular organelles. This is the choice family of dyes for monitoring changes in

plasma membrane potential in vitro [76].

When compared with ArcLight and FluoVolt, PMPI has the advantage of a significantly

greater sensitivity to our nsPEF conditions. Moreover, it requires a simple dilution (0.5 µL/

mL) into the imaging solution and a 30 minute incubation period prior to use. Finally, PMPI

has been relatively well tested and validated as an appropriate tool for the optical

measurement of transmembrane potential, even when compared to patch-clamp

electrophysiology [77–80]. Taken together, this voltage-sensitive dye was selected to

investigate nsPEF effects on Vm in the experiments to be presented.

Page 39: Investigating the role of voltage-gated ion channels in

38

1.15 Image acquisition

Images were acquired using MetaFluor software (Molecular Devices, USA) in Limoges

and with MetaMorph software (Molecular Devices, USA) in Ljubljana. Prior to

experimentation the growth media was replaced with Live cell imaging solution (Molecular

probes-A14291DJ) with the addition of 20 % D - glucose at a concentration of 1.0 mg/mL.

Images were acquired every 30 seconds for experiments using PMPI, and every 3 seconds for

those using YO-PRO. Where electropulsation occurred, a 5 minute baseline of fluorescence

was collected for PMPI experiments, and a 30 second baseline for YO-PRO experiments.

Figure 1.11 – Example of image processing and data extraction. The ROI’s are traced

around the perimeter of each cell membrane in the brightfield image (top left). These ROI’s

are then copied and pasted into the fluorescence image series (top right). Both the raw

fluorescence data (bottom left) and relative fluorescence data (bottom right) is extracted for

statistical analyses.

0

10

20

30

40

0 500 1000 1500 2000Me

mb

ran

e d

ep

ola

riza

tio

n

(Raw

dat

a)

Time (s)

-0.5

0

0.5

1

1.5

0 500 1000 1500 2000Me

mb

ran

e d

ep

ola

riza

tio

n

(ΔF/

Fo)

Time (s)

Page 40: Investigating the role of voltage-gated ion channels in

39

An example of image analysis using HT22 undifferentiated cells in Image Analyst MKII

software (Fig. 1.11). The initial step involves tracing the perimeter of the visible region of each

cell membrane which will be the regions of interest (ROIs) when extracting data. Following

this step, the ROIs were copied from the bright field image and then pasted into the

fluorescence image series. An additional ROI is traced in an area of the fluorescence image

series where no movement is observed. This final ROI will serve to remove the background

noise by subtracting its value from the entire image series. Finally, the raw and relative (ΔF/Fo)

data is extracted and exported for statistical analysis. This was an example where PMPI was

used. YO-PRO experiments were handled in the same manner.

1.16 Statistical analysis

Following image analysis and extraction of data, statistical analyses were performed

using IBM SPSS 19 or SigmaPlot v.11. Multiple types of statistical tests were performed

throughout this thesis, all of which were parametric tests. Therefore, certain assumptions

pertaining to the dataset had to be met for these tests to be properly used and interpreted. The

following assumptions are required for all tests used in this thesis. Additional assumptions are

required depending on the statistical test used. For a review on the more specific assumptions,

refer to [81].

1.16.1 Independence of data

Because the experimental questions looked at differences between groups, the

assumption of independence had to be met. This assumption simply states that the

observations were collected on different subjects. This can be slightly difficult in in vitro

studies since we are often comparing different effects on the same cell line. Therefore, to

ensure the design was as balanced as possible, data was collected and repeated on different

days ensuring that the assumption of independence was met.

Page 41: Investigating the role of voltage-gated ion channels in

40

1.16.2 Normal distribution

Since parametric tests are based off the normal distribution, it is important to ensure

that the data collected within groups is normally distributed. To measure normality, several

measures should be evaluated; including plotting the frequency distribution, evaluating the

shape of the data using values of kurtosis and skewness, analyzing P-P plots to determine if

the data collected was comparable to a theoretical normal distribution, and using the

Kolmogorov-Smirnov (K-S) test which compares the data with a theoretical dataset containing

the same mean and standard deviation.

1.16.3 Homogeneity of variance

The final general assumption to be met involved variability in the dependent variable.

This requires that the variances within conditions are approximately similar. As an example,

suppose we are interested in the electric field threshold for a 10 ns PEF. To meet this

assumption, the variability should be similar for each electric field intensity used. Levene’s test

is used for this purpose, where a significant result suggests that the variance between

conditions is not similar.

1.16.4 Violations of statistical assumptions

Two options are available when the data does not meet the conditions for parametric

testing: Either the data can be transformed using one of many available options, or it can be

removed from the analysis. The drawback to transforming data is that one must also change

any conclusions. For example, if we log transform our data for membrane depolarization, any

conclusions must interpreted in the same way. For the purpose of our results, it seems

unhelpful to make claims about the effects of PEF on the log transformed membrane

depolarization.

Therefore, we are left with the alternative, which is to remove any cases that are

significantly influencing the dataset. For the purpose of analyses outliers were defined as cells

(ROIs) whose fluorescence was beyond 2 standard deviations from the group mean (+/- 2 SD).

Page 42: Investigating the role of voltage-gated ion channels in

41

1.16.5 Additional inclusionary criteria

It is worth noting that prior to testing the assumptions, additional criteria was

necessary for inclusion. In order to be considered for analysis, the cells being measured could

not undergo any excessive movement. This included an automatic exclusion of any cells

entering or exiting the field of view after the experiment commenced. Moreover, cells were

excluded in the event that other debris entered the image series and altered the measured

fluorescence in those ROIs. Finally, a stable baseline was required prior to any experimental

manipulation in order for data from that cell to be further considered.

1.16.6 Definition of threshold

Many of the results to be presented will refer to a threshold. Whether it refers to PMPI

experiments or YO-PRO experiments, the threshold was defined as the minimum electric field

intensity required to produce a response that was significantly different than the control as

determined by the statistical test performed. In the case of PMPI experiments, the threshold

refers to membrane depolarization, whereas in YO-PRO experiments this refers to membrane

permeability. The alpha criteria was set at α = 0.05 for all tests.

1.16.7 Statistical terminology and abbreviations

Each statistical test is associated with an abbreviation. Using an ANOVA, the following

will help elucidate some of the terminology that will be encountered when reading this text.

The first letter, F, is used to represent the ANOVA. If a t-test is used, a t will replace the F; and

in the case of a discriminant analysis, a Λ will take its place. In brackets, 2 numbers are seen

and these represent the degrees of freedom (between, within groups). Between groups degrees

of freedom (df) are calculated by subtracting 1 from the total number of groups tested. This

would indicate that 3 groups were tested in the example below. Within groups df relates more

to the sample size, where the number of groups is subtracted from the total sample size.

Following the same example, this would indicate that the total sample size was 30. The number

Page 43: Investigating the role of voltage-gated ion channels in

42

14.20 is the obtained F-value, which is compared to a critical F-value and is dependent on the

degrees of freedom and the threshold for statistical significance.

F (2, 27) = 14.20, p<0.01, Ω2 = 0.80

The p-value is a measure of statistical significance. If the F-value obtained is greater

than the critical F-value, the p-value will represent a statistically significant difference between

the groups tested. This value is traditionally set to either 0.05 or 0.01, which suggests that

there is a 5 % or 1 % probability that the measured significance is false. The final number is

the effect size Ω2. This value represents the proportion of change in your dependent variable

due to the experimental manipulations of the independent variable. The value 0.80 means

that 80 percent of the variability in your dependent variable is due to the experimental

manipulation. Additional abbreviations will be encountered for parameters such as; sample

size (n), average (X), and standard error (SE).

1.17 Objectives and hypotheses for this Thesis

As a general overview of this chapter, whether the cell or the intracellular organelles

are targeted, ultimately the membranes surrounding them are influenced using pulsed electric

fields. Depending on the desired effect, pulse durations can be used which are greater than or

less than the membrane charging constant. When pulse durations are greater than the

charging time, we observe effects at the level of the plasma membrane. Conversely, shorter

pulse durations have the ability to penetrate the plasma membrane and influence intracellular

organelles.

The research presented in this thesis investigated pulse durations ranging from 10

nanoseconds up to 10 milliseconds on several cell lines. The major point of interest will be to

study the potential for these fields to act on transmembrane proteins, specifically ion channels.

Three experimental chapters will be presented. Because our lab specializes in nsPEF exposure,

the bulk of the data will be centered on these effects. The longer pulse durations were used in

Page 44: Investigating the role of voltage-gated ion channels in

43

the final experimental chapter, which was a collaborative venture between our lab and the

University of Ljubljana.

The first chapter presents a statistical perspective on the results of nsPEF-induced

membrane depolarization. Here we examined experiments on 500 cells over 3 years and

attempted to use factors common in electrophysical models, such as cell size, shape,

orientation to the electric field, and density of cells exposed to determine how well these

parameters could predict the observed depolarization response.

The second chapter investigates the potential for nsPEF to modulate ion channel

activity in U87 cells. Here, a vast number of broad and specific modulators of a variety of ion

channels were used to determine whether they could influence the depolarizing response

following electric field exposure. The null hypothesis being, if the depolarizing response to

these fields was predominately due to electroporation of the plasma membrane, blocking

individual ion channels should not significantly alter the effect.

Finally, knowledge gained from these chapters was applied to several cell lines in a

collaborative project with the University of Ljubljana. Both excitable and non-excitable cell

lines were selected to determine effects of pulse durations ranging from 10 ns to 10 ms on

membrane depolarization. Membrane permeability was an additional parameter explored for

these experiments. The goal for this chapter was to determine whether an electric field

intensity could be found that was able to permeabilize the cells without significantly altering

the transmembrane potential. The application of such a pulse would be to permeabilize cells

targeted during electrochemotherapy treatment, while minimizing pain associated with

electrically exciting surrounding nervous tissue. A manuscript has been submitted for this

collaboration and is currently under review. The results presented in this thesis will include

only those that I directly contributed to.

Page 45: Investigating the role of voltage-gated ion channels in

44

Chapter 2:

A comparative analysis of the

theoretical and experimental

interactions of PEF with cells in vitro

Page 46: Investigating the role of voltage-gated ion channels in

45

2.1 Introduction

Despite decades of research into the effects of PEF on biological systems, the

mechanisms by which electric fields interact with cells are not fully understood. The present

consensus relies on the combination of mathematical models and computer simulation to

describe and predict experimental data.

A pioneer in this field of research, H. P. Schwan, was the first to mathematically

describe the interactions between electric fields and biological systems in the late 1950’s [82].

This relationship is summarized in Fig. 2.1 below, where the induced transmembrane

potential, (Vm), is expected to be proportional to the applied electric field (E) and the cell radius

(R). The cosine function predicts that Vm will be non-uniformly affected along the cell

membrane, such that the greatest effects would be predicted at areas of the membrane

adjacent to the electrodes. Moreover, it suggests that when two cells are exposed to an electric

field of equal intensity, the greatest effects on Vm will be observed in the larger cell.

𝑉𝑚 =3

2 𝐸𝑅 𝑐𝑜𝑠𝜃

Figure 2.1 – Schwan equation (steady - state). This steady-state equation predicts the

induced transmembrane potential (Vm) on a cell with a radius (R) following the application of

an electric field with an intensity (E). The cosine function (θ) suggests the effect will be

influenced by the angle at which it interacts with the membrane.

An important issue to consider with the steady-state Schwan equation is that it does

not consider the time required to superimpose an external electric field on the cell’s resting

membrane potential. This has become especially important as pulse durations have become

Page 47: Investigating the role of voltage-gated ion channels in

46

shorter than the time constant for charging the plasma membrane. For that to be considered,

a modification of Schwan’s equation incorporates the pulse duration (t) and the membrane

charging time (τm) (Eq 2.1).

𝑉𝑚 =3

2 𝐸𝑅 𝑐𝑜𝑠𝜃 (1 − 𝑒

(−𝑡𝜏𝑚

))

Equation 2.1 - Schwan equation (first - order). Extension of the Schwan equation that

incorporates time (t), and the membrane charging constant (τm).

One limitation with the Schwan equation is that the more a cell deviates from a

spherical shape, the less it becomes applicable. With that said, regardless of the shape of the

cell, experimental data has demonstrated repeatedly when fluorescent markers with high

temporal resolution are used, the greatest effects are observed at membrane regions adjacent

to the electrodes [83–86].

An excellent example of this was published recently, when a genetically encoded

calcium indicator, GCamP, was used in experiments using nsPEF [87]. In Fig. 2.2 the authors

delivered 100 x 10 ns pulses at either 10 Hz (top series) or 1 Hz (bottom series) with an electric

field intensity of 44 kV/cm. In both conditions, although the time dynamics differ, a wave of

calcium can be seen entering the cell moving along the electric field gradient from the anode

to the cathode.

Sometimes, however, we find unanticipated results in the literature. Theory would

predict [88] that, under the same conditions, the cell with a larger radius should be more

influenced by the electric field than the smaller one. Yet we find results that show a positive

relationship [88], an inverse [89], or no relationship [90, 91] between the size of a cell and the

observed effects.

Page 48: Investigating the role of voltage-gated ion channels in

47

Figure 2.2 – Calcium wave entering the cell from the anodic pole following

nsPEF. The top series shows results following 100 x 10 nsPEF, delivered at 100 Hz and an

electric field intensity of 44 kV/cm. The bottom exposed cells to the same parameters except

the pulses were delivered at 1 Hz. The numbers above the images are measure of time in

seconds. Figure adapted from [87].

It is possible that there are some methodological differences that could shed some light

on this discrepancy. For example, there were different cell types used in those experiments,

which included yeast, rodent and human cell lines. All cells exhibit fairly characteristic growth

patterns and rates, specific to their cell type. Since the density of cells being exposed is

expected to play a significant role in the observed effect [86, 92], perhaps this factor could

have contributed to the variation in results when considering the size of the cell. Despite the

models describing the influence of cell density on PEF effects, I was unable to find any

published articles that attempted to experimentally account for this parameter.

The objective for this section will be to determine how factors such as cell size, shape,

orientation and density can predict membrane depolarization in U87 cells exposed to a single

10 ns PEF. Although analytical and numerical models for these parameters have been

described [85, 86, 92], (Fig. 2.3) there have been no studies which have tried to experimentally

determine how all of these factors simultaneously influenced the induced transmembrane

Page 49: Investigating the role of voltage-gated ion channels in

48

potential. For these analyses, additional factors such as temperature, medium conductivity,

electric field intensity, pulse shape and duration have all been held constant to reduce as many

extraneous variables as possible. One would expect that a significant contribution to the

induced transmembrane potential following a PEF will be accounted for by variables such as;

cell size, shape, the angle at which the PEF interacts with the cells major axis, and the density

of cells in the field of view.

Figure 2.3 – Theoretical interaction of an electric field and a cell membrane.

Factors such as cell size (top left), shape (top right), orientation to the electric field (bottom

left), and density of cells (bottom right) will influence the observed response following PEF.

2.2 Methods

2.2.1 Determining size, shape and orientation of cells to PEF

2.2.1.1 Calibration of images

In order to measure the size of a cell, images had to be calibrated. This was

accomplished using Image J software along with a calibration slide. The slide contained

15 micron beads, and when loaded into Image J, a line was traced across the diameter of the

bead. A measurement of the line provided a value in pixels, and since the diameter of the bead

Page 50: Investigating the role of voltage-gated ion channels in

49

was known, we obtained a value of pixels / µm. This value was then applied to all subsequent

images.

2.2.1.2 Determining size of the cell

The size or area of a given cell was determined using the same protocol as in 2.4.3,

where a line was traced around the visible perimeter of the plasma membrane. This procedure

was repeated for all cells in the field of view, and then added to the ROI manager in Image J.

From here the measurements for size (µm2) were extracted and loaded into SPSS for analysis.

2.2.1.3 Determining shape of a cell

The cells used in these analyses were attached to the glass dish, which means that they

each varied in shape to a certain degree. Image J offers a measurement of circularity, which is

described in equation 2.2 below. This provides a unitless value with a range from 0 to 1, with

1 indicating a perfect circle.

𝐶𝑖𝑟𝑐𝑢𝑙𝑎𝑟𝑖𝑡𝑦 = 4𝜋 𝑥 𝐴𝑟𝑒𝑎

𝑃𝑒𝑟𝑖𝑚𝑒𝑡𝑒𝑟2

Equation 2.2 - Determining the shape of a cell. A predetermined parameter provided

by image J software providing a measure of sphericity. Unitless values are computed from 0 –

1, where 1 indicates a perfect circle.

2.2.1.4 Determining the angle of the cell with respect to the applied electric field

The ROI manager in Image J also provides a measurement of angle. This is a

calculation that initially measures the long and short axes of the plasma membrane, and then

measures the angle in degrees between the x axis of the image and the cell’s long axis. Once

this data was extracted, a correction was made for the positioning of the electrodes.

2.2.1.5 Determining the density of cells in a given image

After each cell had been traced a final ROI was traced around the entire field of view,

which was also added to the ROI manager. The density of cells per unit area (cells / µm2) was

Page 51: Investigating the role of voltage-gated ion channels in

50

then calculated simply by dividing the number of cells in the field of view by the area of the

image.

2.2.2 Statistical analyses

The statistical tests used in this chapter were based on the assumptions of a parametric

dataset. As such, various criteria had to be met prior to performing these tests. These

assumptions have been outlined in the general methods section of the introductory chapter

and are detailed in [81].

2.3 Results

From a large sample size of 500 U87 cells collected over a multitude of experiments

and over a three year period, variables such as cell size (µm2), circularity of cell (1 indicating a

perfect sphere), angle (degrees) relative to the direction of the applied electric field, and

density of cells in the field of view (cells per µm2) were extracted. These factors were

subsequently entered as predictors into a linear multiple regression. The outcome variable was

relative change in membrane potential, which was measured using PMPI.

Two variables were able to account for a small proportion of the change in Vm following

nsPEF. Circularity (n = 500, X = 0.78, SD = 0.10, p < 0.01) and density (n = 500, X = 0.0015,

SD = 0.0008, p = 0.02) entered the statistically significant model [F(2 , 497) = 6.65, p < 0.01,

Ω2 = 0.02]. The relationship can be represented by (Eq. 2.3). This equation, as would be

expected, highlights a positive relationship with circularity and a negative relationship with

cell density. In other words, the greatest change in Vm post nsPEF would be expected under

conditions where fewer cells were present and when those cells were more spherical than

irregularly shaped.

Page 52: Investigating the role of voltage-gated ion channels in

51

𝑉𝑚 = 88.4𝑥 − 9157.9𝑦 + 54.9

Equation 2.3 – Equation describing relationship between cell shape and density

with PEF-induced membrane depolarization. Here, Vm is measured as relative change

in PMPI fluorescence; x refers to the shape (circularity); and y refers to density.

It is interesting that the statistically significant model was only able to account for

approximately 2 % of the change in Vm following nsPEF exposure. In other words, that

majority of the variability in membrane depolarization is unaccounted for using the factors

entered into the regression. Moreover, factors such as cell size and orientation to the electric

field failed to significantly contribute to the model.

After carefully looking through the data, something became apparent that wasn’t

noticed during the initial analysis. When considering the baseline fluorescence, there was a

clear grouping of data points (Fig 2.4). To test if this represented two statistically different

groups, a sequence of tests was performed. To begin, a cluster analysis was performed on the

baseline fluorescence data. Consistent with our observations, the results from the cluster

analysis divided the data into 2 groups, where 465 cells were in the first group and the other

35 cells were in the second group.

Using the groups generated from the cluster analysis, a discriminant function analysis

was performed to determine if these groups were truly statistically different from one another.

The results from this test revealed a discriminant function that explained 92% of the variance

between groups. This function significantly differentiated the baseline fluorescence data [Λ =

0.15, χ2(1) = 946.82, p < 0.01].

Page 53: Investigating the role of voltage-gated ion channels in

52

Figure 2.4 - Investigating pattern in baseline fluorescence. Two distinct clusters are

easily distinguishable when examining baseline fluorescence values. X (average of first 5

minutes pre-pulse) and y (first 2.5 minutes post-pulse) values are given in raw fluorescence

values.

These results suggest that there were two populations of cells that were significantly

different from one another despite having identical experimental protocols. Moreover, these

cells weren’t clustered temporally, meaning they represent two distinct populations that have

been collected over several years. What could be different about these cells then? Referring to

table 2.1 and 2.2 below, you will find descriptive statistics for the two groups along with a

correlation table which describes their relationship with the factors used in the multiple

regressions.

Cluster 1 Cluster 2

Page 54: Investigating the role of voltage-gated ion channels in

53

Sample size (n)

Minimum Maximum Mean Standard deviation

Cluster 1

Area (µm2) 465 15.09 538.62 87.05 58.56

Baseline fluorescence

465 12.55 465.64 88.48 84.38

Circularity 465 0.36 0.96 0.77 0.10

Angle (degrees)

465 0.25 179.76 95.66 51.50

Density (#/µm2)

465 9.70 10-5 2. 10-3 1.54 10-3 8.54 10-4

Cluster 2

Area (µm2) 35 33.33 363.48 126.24 74.08

Baseline fluorescence

35 543.81 1237.91 958.54 172.66

Circularity 35 0.65 0.90 0.80 0.08

Angle (degrees)

35 0.13 177.49 92.32 46.94

Density (#/µm2)

35 2.91 10-4 2.82 10-3 1.48 10-3 4.56 10-4

Table 2.1 - Descriptive statistics for the two populations of cells grouped by

baseline fluorescence. For each parameter investigated, this table provides the sample size

(n), minimum, maximum, and mean values, as well as the standard deviation for both clusters

of cells.

Page 55: Investigating the role of voltage-gated ion channels in

54

Fluorescence change

Area (µm2)

Fluorescence baseline

Circularity Angle (degrees)

Density (#/µm2)

Cluster 1 0.106 0.424 0.123 0.029 -0.107

Cluster 2 -0.160 -0.446 0.069 0.238 -0.072

Table 2.2 - Pearson correlation table from the multiple regression test divided

into clusters. Values in bold indicate a statistically significant correlation between those

values and the change in fluorescence using an alpha of 0.05.

Interpretation of the above two tables provides insight into the difference between

groups. Beginning with the descriptive statistics (Table 2.1), aside from the different baseline

fluorescence values, the average values from all other factors are essentially identical.

Important differences are seen, however, when we refer to the correlation table (Table 2.2).

Despite being the same cells treated in the exact same way, the first cluster is comprised of

cells that are depolarized more as the size of the cell increases; which is contrasted in the

second cluster, where depolarization is more pronounced as the size of the cell decreases.

The same pattern follows when considering the baseline values. Keeping in mind that

the baseline values are representative of the resting membrane potential, cells in cluster 1 have

a tendency to be more influenced by exposure to the nsPEF as the baseline fluorescence

increases (less negative resting membrane potential); whereas cells in cluster 2 tend to be

more influenced as the baseline fluorescence decreases (more negative resting membrane

potential).

2.4 Discussion

Since the 1950’s, electrophysical models have described the interaction between an

applied electric field and a biological membrane. These have been extremely useful in

designing experiments and interpreting results. Yet, despite multiple discrepancies between

Page 56: Investigating the role of voltage-gated ion channels in

55

theoretical predictions and experimental data in the literature, the established theory has not

been challenged.

Of course these models are based off well established physical laws of

electromagnetism described eloquently by Maxwell in the mid 1800’s; however, application of

these theories into biology tends to require that a cell is reduced from a complex living thing

to a capacitor floating in conductive solution.

Molecular dynamics (MD) simulations have been a technological breakthrough that

allow us to investigate electric field effects on the individual molecules making up a cell. Yet

despite this incredible tool, the programs that determine the output are all designed based off

the same fundamental theories. One advantage of MD is modeling a dynamic structure as

opposed to the static electrophysical resistor-capacitor model. The cost of this, however, is the

requirement for tremendous computational power limiting our search to a small fraction of

the cell membrane and for a very short period of time.

The goal of this chapter was to investigate how theoretical models hold up against a

large experimental dataset. Membrane potential was selected as the outcome variable because

according to the literature, the induced transmembrane potential is the first step in a series of

biological events following electropulsation [17, 18].

The results from this section suggest that under these conditions, there

appears to be additional variables that are contributing to the nsPEF-induced

membrane depolarization. Factors such as cell size, shape, orientation to the electric field

and the density of cells being exposed failed to account for more than 90 % of variability in

membrane depolarization following nsPEF exposure.

So, what other factors could account for this variability? One possibility could include

ion channels. This idea was first explored almost 30 years ago [93], where microsecond pulses

were found to influence the activity of Na+/K+ pump in human erythrocytes. Later, evidence

was presented suggesting ms-PEF may induce electro-conformational damage to voltage-gate

Page 57: Investigating the role of voltage-gated ion channels in

56

channels [94]. Despite these findings, researching the potential role of ion channels in PEF-

effects has just recently began to resurge. Several studies have now shown that voltage-gated

ion channels can be activated by nsPEF [63–65].

Directly related to the activity of voltage-gated channels, the cell cycle could provide

additional clues regarding variability in data from PEF experiments. Cells constantly oscillate

through periods of de- and hyperpolarization, associated with changes in ion channel

conductivity, that have been directly correlated with specific phases of the cell cycle [95–97].

When we consider results [89] demonstrating that membrane permeability was

significantly greater following PEF exposure of cells in the S – M phase of the cell cycle

compared to those in the G1 phase, it may not be spurious that we found two statistically

distinct groups of cells clustered by baseline fluorescence (resting membrane potential).

Furthermore, the 90/10 split between clusters is in line with the proportion of time primary

carcinomas [98], and specifically U87 cells [99], spend in G1 phase versus S, G2 and M phases.

Even If we assume that our clusters are representative of groups of cells divided into G1 and S

– M phases, several key differences make it is difficult to compare our results to those in [89];

namely different pulse durations (µs vs ns), measured outcomes (permeability vs membrane

potential), and cell type used (yeast vs human cells).

With that said, more attention should be given to expanding our understanding of

electric field – cell membrane interactions. This is most important when we are looking at cell

effects over longer periods of time. Combining the results presented here, along with some of

the recent research discussed above, the influence of PEFs on transmembrane ion channels

and perhaps even the role of cell cycle effects seem like good candidates for future studies.

Page 58: Investigating the role of voltage-gated ion channels in

57

Chapter 3:

Nanosecond pulsed electric fields

depolarize transmembrane poten-

tial via voltage-gated K+, Ca2+ and

TRPM8 channels in U87

glioblastoma cells.

(Based on published manuscript)

Page 59: Investigating the role of voltage-gated ion channels in

58

3.1 Introductory Remarks

The following chapter is based on the manuscript [100]. It is worth noting that

additional experiments have been included here that were not present in the manuscript.

When considering the effects of PEF on complex living cells, the most fundamental process to

begin our investigations would be the transmembrane potential (Vm). The reason it is

imperative to understand the interaction of an electric field with the cell’s resting membrane

potential is because the superposition of the applied electric field on the cells Vm is considered

to be the initial factor influencing cell behavior [17, 18] .

nsPEFs have a variety of applications in the biomedical and biotechnology industries.

Cancer treatment has been at the forefront of investigations thus far as nsPEFs permeabilize

cellular and intracellular membranes leading to apoptosis and necrosis. nsPEFs may also

influence ion channel gating and have the potential to modulate cell physiology without

poration of the membrane. This phenomenon was explored using live cell imaging and a

sensitive fluorescent probe of transmembrane voltage in the human glioblastoma cell line, U87

MG, known to express a number of voltage-gated ion channels. The specific ion channels

involved in the nsPEF response were screened using a membrane potential imaging approach

and a combination of pharmacological antagonists and ion substitutions. It was found that a

single 10ns pulsed electric field of 34 kV/cm depolarizes the transmembrane potential of cells

by acting on specific voltage-sensitive ion channels; namely the voltage and Ca2+ gated BK

potassium channel, L- and T-type calcium channels, and the TRPM8 transient receptor

potential channel.

3.2 Introduction

Nanosecond pulsed electric fields (nsPEFs) have been exploited in applications such

as cancer therapy [66, 101, 102], and have demonstrated promise in improving gene delivery

when combined with conventional pulses [103, 104]. In vitro experiments [27, 69, 105–107]

Page 60: Investigating the role of voltage-gated ion channels in

59

along with support from mathematical modeling [32, 62, 108] have demonstrated that nsPEFs

cause changes in cell physiology by permeabilization of the plasma membrane and

intracellular organelles. Results have also been demonstrated in vivo with considerable nsPEF

effects on tumor growth [55] and vascular perfusion [109]. A prominent mechanistic working

model developed from MD simulations suggests that nsPEF permeabilization is mediated by

the reorganization of water molecules at the lipid membrane interface by the electric field-

induced orientation of their dipoles, leading to the formation of pores [110].

Few experimental studies or simulations have considered the potential role of

transmembrane pore-forming protein complexes and channels in nsPEF effects. A wide

variety of such ion channels exist in the plasma membrane that can influence membrane

permeability to ions and small molecules and can be gated by a wide range of stimuli ranging

from ligands (ionotropic receptors), voltage (voltage-gated ion channels), mechanical

perturbation (mechanosensitive ion channels) or changes in temperature (transient receptor

potential channels (TRPs)).

In vitro studies have primarily investigated the permeabilization of the plasma

membrane by nsPEF using the translocation of small membrane-impermeant fluorescent

molecules, fluorescent ion sensors, or electrophysiology. The latter, more direct approach of

measuring plasma membrane impedance using the patch-clamp technique was used to probe

membrane resistance following the application of nsPEFs (60 ns, 12 kV/cm), revealing a

threefold decrease in membrane resistance 80 - 120 seconds following pulse delivery with a

gradual recovery over 15 minutes [105]. Several ion channel pore blockers and ion

substitutions were used in these experiments to examine the selectivity and specificity of the

membrane permeability to various ionic species. Although this study did support MD theory

of membrane permeabilization, the authors were careful to differentiate permeabilization and

poration, and concluded that there was not sufficient evidence to confirm the creation of

aqueous pores.

Page 61: Investigating the role of voltage-gated ion channels in

60

A subsequent study examined the size and lifetime of nanopores at the level of the

plasma membrane. These experiments compared the effects of nsPEF (600 ns, 0.75-10.2

kV/cm) on the uptake of small fluorescent molecules (YO-PRO, propidium iodide PI) vs the

uptake of Thallium (Tl+) [111]. This investigation followed previous studies [69, 112] which

demonstrated that more intense electric field strengths were required to observe PI uptake

versus YO-PRO uptake. It was concluded that the larger PI molecules were occluded from

pores where the smaller YO-PRO molecules passed through. Because of its small size (< 1 nm),

Tl+ uptake showed the presence of even smaller nanopores from which YO-PRO and PI would

be excluded. The results from this study, combined with previous findings, suggested that

nsPEF exposure led to pores smaller than 1 nanometer.

Some studies have considered the possibility that such short pulsed electric fields could

influence ion channel behavior. This alternative to electroporation was first explored in bovine

chromaffin cells using a pulse protocol of 5 ns, 50 kV/cm [64]. By using several

pharmacological inhibitors of calcium and sodium channels, it was shown that a combination

of L- N- and P/Q-type calcium channel blockers abolished the nsPEF-induced influx of

calcium into cells, suggesting the involvement of voltage-gated calcium channels.

In a separate study, the effects of PEF’s was examined on several cell lines [63] using a

pulse duration of 500 picoseconds at 190 kV/cm, with a range of pulses up to 100 at a

frequency of 200 Hz or 1 kHz. When applying these pulse parameters to CHO cells, which

served as a negative control due to the lack of voltage-gated channels, no effects were observed

while measuring [Ca2+]i with Fura-2. Yet when tested on cell lines known to contain voltage-

gated channels, such as GH3 and NG108, the same protocol resulted in a significant increase

in [Ca2+]i. To demonstrate this effect was not due to electroporation of the plasma membrane,

voltage-gated calcium channel blockers, verapamil (L-type) and w-Conotoxin (N-type) were

utilized. While these pharmacological inhibitors could partially block the calcium response

when used separately, as a cocktail they could block 85 - 100 % of the observed PEF response.

Page 62: Investigating the role of voltage-gated ion channels in

61

These results imply that the calcium increase caused by ultrashort pulsed electric fields was

mediated by voltage-gated calcium channels.

The distinction between direct permeabilization effects on the plasma membrane and

changes in ion permeability mediated by voltage-gated ion channels is important, given that

many cancers are known to express voltage-gated ion channels [113, 114]. We therefore chose

to investigate the influence of nsPEFs on the human glioblastoma cell line (U87), as it

expresses many types of ion channels that are characterized at the level of gene expression

[115] and in detailed electrophysiological recordings [116, 117]. Our working hypothesis was

that if nsPEF-induced membrane permeabilization was due to electroporation and the

formation of nanopores, the response would not be inhibited by specific pharmacological

inhibitors of ion channels. We tested this hypothesis using a fluorescent indicator of

transmembrane potential to screen a wide range of ion channel inhibitors for their ability to

inhibit a plasma membrane depolarization response caused by a single, 10 ns pulsed electric

field of 34 kV/cm.

3.3 Materials and Methods

3.3.1 Pharmacological manipulation of ion channel activity

Multiple ion channel agonists/antagonists were used to monitor their effects on

nsPEF-induced membrane depolarization. These drugs were selected based on their ability to

block channels known to be present in this cell line [97, 116–121]. All ion channel modulators

were purchased from Sigma Aldrich and are summarized by activity in table 3.1 below.

Page 63: Investigating the role of voltage-gated ion channels in

62

Product Concentration Mechanism of action

Incubation period

Calcium channel modulators

A784168 10 µM TRPV1 blocker 15 minutes

AMTB Hydrate 50 µM TRPM8 channel blocker

15 minutes

BAPTA-AM 10 µM, 30 µM IC Ca2+ chelator 45 minutes

Cyclopiazonic acid 1 µM Ca-ATPase blocker/depletes ER

10 minutes

Ethylene glycol tetraacetic acid (EGTA)

1.8 mM (replaced CaCl2)

EC Ca2+ chelator 15 minutes

Gadolinium 10 µM Broad calcium and TRP channel antagonist

15 minutes

HC-030031 3 µM to 300 µM TRPA1 blocker 15 minutes

Ionomycin 10µM Ca2+ ionophore Added prior to imaging

Lanthanum chloride (LaCl3)

100 µM Broad calcium channel antagonist

15 minutes

M8-B 2 nM to 50 µM TRPM8 channel blocker

15 minutes

Mibefradil 20 µM T-type VGCC blocker 15 minutes

Nifedipine 10 µM L-type VGCC blocker 15 minutes

Ruthenium Red 100 µM Broad calcium channel antagonist

15 minutes

2-APB 100 µM TRPC blocker 15 minutes

Potassium channel modulators

Barium chloride (BaCl2)

2.5 mM (replaced KCl) KIR channel blocker 15 minutes

Clotrimazole 30 µM IK channel blocker 15 minutes

Glibenclamide 50 µM ATP-dependent K+ channel blocker

15 minutes

Iberiotoxin 10 nM BK channel blocker 15 minutes

Page 64: Investigating the role of voltage-gated ion channels in

63

Paxilline 100 µM BK channel blocker 15 minutes

Penitrem A 100 nM to 10 µM BK channel blocker 15 minutes

Phloretin 30 – 1000 µM BK activator Added prior to imaging

Tetraethylammonium chloride (TEA)

5 mM to 100 mM Voltage-gated K+ channel blocker

15 minutes

Sodium channel modulators

Amiloride HCl 200 µM Na+ channel blocker; ASIC blocker; mechanogated ion channel blocker

15 minutes

Tetrodotoxin (TTX) 1 µM Voltage-gated Na+ channel blocker

15 minutes

Choline Chloride 140 mM Used to replace sodium in imaging media

15 minutes

N-methyl-D-glucamine (NMDG)

140 mM Used to replace sodium in imaging media

15 minutes

Table 3.1 - List of pharmacological agents used throughout this investigation.

Product names are given along with the concentrations used and their mechanism of action.

For clarification of abbreviations: TRP = transient receptor potential, IC = intracellular, EC =

extracellular, CaMKII = Calcium/calmodulin-dependent protein kinase II, VGCC = voltage-

gated calcium channel, KIR = inward-rectifying potassium channel, IK = intermediate

conductance calcium-dependent potassium channel, BK = big conductance calcium-

dependent potassium channel, ASIC = acid-sensing ion channel.

3.3.2 Calibration of PMPI voltage-dye

In order to obtain information regarding membrane potential from fluorescence data

the fluorophore (PMPI) needed to be calibrated. Cells were incubated in the same method as

above; however, several imaging solutions were made where the concentration of NaCl was

replaced by increasing amounts of KCl (2.5 mM, 25 mM, 50 mM, 100 mM, and 140 mM). An

Page 65: Investigating the role of voltage-gated ion channels in

64

estimate of membrane potential was then computed by applying the Goldmann equation and

comparing the fluorescence data at various K+ concentrations with the calculated membrane

potential at those concentrations. Although semi-quantitative in nature, this method allowed

for a general estimate of the changes in membrane potential over time following exposure to

nsPEF.

The results from the calibration were compared to separate results calibrating PMPI

using patch-clamp electrophysiology (section 3.4.1). Here, cells were exposed to a voltage-

ramp while PMPI fluorescence was measured. With these results, we were able to compare

both quantitative and semi-quantitative measures of transmembrane potential and use this

data to estimate the magnitude of depolarization following electric field exposure.

3.3.3 Statistical analyses

Statistical analysis was performed using IBM SPSS 19. Assumptions were met for

analysis using parametric tests. More detail on the individual statistical analyses can be found

in the general methods section of the introductory chapter, in addition to [81].

3.4 Results

3.4.1 Calibration of PMPI

The first step in our investigation was to calibrate PMPI. Although semi-quantitative

in nature, using different calibration methods it should be possible to determine a change in

transmembrane potential in response to a stimulus. This is different than knowing the

absolute value of the transmembrane potential at a given moment; however, knowing how

much a given manipulation is altering Vm is still useful for interpreting effects. PMPI

calibration was performed using potassium [K+] gradients and additionally using patch-clamp

electrophysiology.

Page 66: Investigating the role of voltage-gated ion channels in

65

3.4.1.1 Potassium calibration

Calibration of the dye using K+ was achieved by monitoring the change in relative

fluorescence over time using varying concentrations of extracellular K+. For a period of 30

minutes, imaging medium was exchanged every 5 minutes with [K+] from 2.5 mM to 140 mM

(Fig. 3.1). Maximum depolarization of the transmembrane potential occurred following the

addition of 100 mM [K+].

Figure 3.1 - Calibration of PMPI by varying extracellular K+. Relative fluorescence

changes for PMPI were measured over time in response to increases in extracellular [K+].

Potassium concentration was gradually increased from 2.5 mM to 140 mM with a step

occurring every 5 minutes Results are the average from 5 experiments. Error bars represent

+/- 2 SE

Because we were most interested in the change in relative fluorescence at different

concentrations of extracellular [K+], a linear regression was performed to determine the

relationship between variables. It should be noted that maximum depolarization occurred at

100 mM; therefore, the 140 mM data was not entered into the regression because no useful

information could be obtained. The relationship between relative fluorescence and

Page 67: Investigating the role of voltage-gated ion channels in

66

extracellular [K+] can be summarized using Eq. 3.1 below in which 96 % of the variability can

be explained by the relationship between variables.

𝑦 = 0.0993𝑥 − 0.479 , 𝑅2 = 0.96

Equation 3.1 – Equation describing relationship between extracellular

potassium concentration and membrane depolarization between 0 and 100 mM

[K+]. A linear regression was performed to find a relationship between relative fluorescence

and the extracellular [K+].

3.4.1.2 Calibration using electrophysiology

The next step involved a similar approach as above, using patch-clamp

electrophysiology. Cells (n = 4) were exposed to a voltage-step ranging from -90 mV to +10 mV

for five minutes and the relative fluorescence was recorded. The results obtained from these

experiments were best fit using a 2nd order polynomial expression, which is summarized in

equation 3.2 below, where the experimental manipulation accounted for 99% of the variability.

This equation suggests that every increase in 1 relative fluorescent unit is associated with a

membrane depolarization equivalent to 22.7 mV.

𝑦 = 0.0052𝑥2 − 0.1198𝑥 + 1.0402 , 𝑅2 = 0.99

Equation 3.2 – Equation describing relationship between membrane potential

and PMPI fluorescence. A 2nd order polynomial equation was best able to explain the

relationship between the transmembrane potential (mV) and the relative fluorescence

measured.

PMPI calibration using two methods should allow us to generate a semi-quantitative

estimate regarding the PEF effects on Vm. Table 3.2 below provides a comparison of the

relative fluorescence changes using K+ calibration and electrophysiology. The last column

Page 68: Investigating the role of voltage-gated ion channels in

67

provides an estimation, using the Goldman-Hodgkin-Katz (GHK) equation (Eq 3.3), of the

transmembrane potential changes associated with alterations of the extracellular potassium

concentration. Results from both electrophysiology and GHK estimation of the Vm are seen in

columns 2 & 4. Relative fluorescence changes using PMPI under both calibration methods

result in very similar (+/- 4 mV) changes in Vm.

𝑉𝑚 =𝑅𝑇

𝐹 ln (

𝑝𝑘 [𝐾+ ]𝑖 + 𝑝𝑁𝑎 [𝑁𝑎+ ]𝑖 + 𝑝𝐶𝑙 [𝐶𝑙− ]𝑖

𝑝𝑘 [𝐾+ ]𝑜 + 𝑝𝑁𝑎 [𝑁𝑎+ ]𝑜 + 𝑝𝐶𝑙 [𝐶𝑙− ]𝑜)

Equation 3.3 – Goldmann equation. Describes the predicted membrane potential based

on the concentrations of ions inside [X]i and outside [X]o of the cell, where X is Na+, K+, or Cl-

, in addition to factors such as membrane permeability (p) of a give ion species, The gas

constant (R), Faraday’s constant (F) and the temperature (T).

Relative fluorescence

(ΔF/Fo)

ΔVm (mV) Ephys

[K+]o ΔVm (mV)

GHK estimation

1 22.7 14.9 18

2 29.3 25.0 26

3 34.1 35.0 32

4 38.0 45.1 37

5 41.4 55.2 42

6 44.5 65.2 45

7 47.3 75.3 48

8 49.9 85.4 51

Table 3.2 - Comparison between calibration methods for PMPI. This table

summarizes the membrane potential change following nsPEF delivery for both the

electrophysiology and potassium calibration.

3.4.2 Determining the threshold of electric field intensity required for nsPEF-induced membrane depolarization.

In order to determine the threshold for membrane depolarization, electric field

intensity was varied (16.5 kV/cm, 22 kV/cm, 34 kV/cm, and 44 kV/cm). When comparing the

Page 69: Investigating the role of voltage-gated ion channels in

68

group means 2.5 minutes post pulse, significant group differences were observed. Applied

electric field strengths of 16.5 kV/cm or 22 kV/cm were not significantly different than the no

pulse conditions, although the latter was qualitatively different. Both the 34 kV/cm and 44

kV/cm applied field strengths were significantly different than control conditions (Fig. 3.2).

In each case, membrane depolarization caused by a single 10 ns pulse was transient,

recovering in 15 - 20 minutes.

Figure 3.2- Determining the electric field threshold for a single nsPEF to

depolarize plasma membrane. An analysis of variance indicated there was a significant

difference in relative fluorescence between groups exposed to different electric field strengths

[F (4, 26) = 27.82, p < 0.01, Ω2 = 0.80]. Post hoc analysis followed using the Games-Howell

method. Significant membrane depolarization was observed when using applied field

strengths of 34 kV/cm [n = 5, X = 2.85, SE = 0.17, p < 0.01, r = 0.99] and 44 kV/cm [n = 5, X

= 2.90, SE = 0.56, p < 0.01, r = 0.97] when compared to control conditions [n = 8, X = 0.01,

SE = 0.02]. When field strength was reduced to either 22 kV/cm [n = 5, X = 0.89, SE = 0.31]

or 16.5 kV/cm [n = 5, X = 0.001, SE = 0.03] no significant effect was observed, though a visible

qualitative response in the 22 kV/cm condition. Pulse deliver occurred at 5 minutes and is

indicated by arrow in figure. Error bars represent +/- 2 SE

Page 70: Investigating the role of voltage-gated ion channels in

69

3.4.3 Fluorescence imaging of plasma membrane depolarization following a single 34 kV/cm nsPEF.

One of the most conspicuous features of the membrane potential dye (PMPI) is its

unequal distribution within the cell. The 2-part dye responds to changes in membrane

potential by entering or exiting the cell. During depolarization, the anionic, charged

component enters the cell resulting in an increase in fluorescence (Fig. 3.3 a, b). This response

is reversed during repolarization as the charged molecules exit the cell and interact with the

quenching agent present in the extracellular space thereby decreasing the fluorescence.

Figure 3.3 - Fluorescence imaging of U87 cells prior to and after delivering a

single nsPEF with an electric field intensity of 34 kV/cm. PMPI is a two-part indicator

of membrane potential consisting of an anionic, charged component and a quenching agent

restricted to the extracellular space. Upon depolarization of the plasma membrane, the

charged molecules enter the cytosol and increase in fluorescence intensity. Pseudo-colored

images are shown for enhanced contrast. a and b depict the fluorescence change pre vs. post

pulse. c is a surface plot montage of the selected region in a framed by the square. The nsPEF

was delivered after five minutes, corresponding to the tenth image of the montage marked by

the arrow. Maximum depolarization can be seen after 60 seconds and individual regions of

Page 71: Investigating the role of voltage-gated ion channels in

70

intense depolarization are observed throughout the remainder of the experiment at different

time points. d is an example taken from the surface plot highlighting the geometry of these

punctate regions where depolarization of membrane potential first occurs in response to the

nsPEF (x and y axes are given in micrometers, while the z axis is in raw fluorescence changes).

3.4.4 Inhibition of nsPEF depolarizing effect with BK channel blockers.

In order to investigate the role of ion channels in the nsPEF depolarization, multiple

ion channel antagonists were tested for their ability to abrogate the response to a single

34 kV/cm nsPEF. If depolarization was mediated by pores, then none of the pharmacological

channel-blockers would be expected to inhibit the response. The following section investigates

the influence of nsPEF on a subtype of voltage-dependent potassium channels.

U87 cells are known to express several potassium channels, many of which are voltage

dependent. Because the observed depolarization was slow, taking place over minutes, non-

voltage-gated channel antagonists were also used. Inhibitors of voltage-and-calcium-

dependent potassium channels were found to inhibit the depolarization of the plasma

membrane caused by a single nsPEF (fig. 3.4 below). When inhibitors of voltage-and-calcium-

dependent potassium channels (BK, SK, IK), Inward rectifying potassium channels (KIR) and

ATP-dependent potassium channels (KATP) were compared with nsPEF-only conditions, there

was a statistically significant difference in their ability to abrogate the nsPEF depolarization

(fig. 3.4 below).

Prior to depolarization, the dye can be seen clustering into small, punctate regions.

Although the spatial distribution of labelling varied in each cell, the clusters accounted for

roughly 15 - 20 percent of the observable plasma membrane, with the small punctate regions

ranging from 0.5 - 1.0 µm in diameter (fig. 3.3 c, d).

Page 72: Investigating the role of voltage-gated ion channels in

71

Post-hoc analysis using Games-Howell method indicated the source of the significant

difference was between the nsPEF only condition and those cells pretreated with 10 μM

Penitrem A and 100 mM TEA. No significant differences were observed between nsPEF

depolarized cells and cells those treated with any of the other inhibitors. Results are presented

in Fig. 3.4 below.

Figure 3.4- Effects of potassium channel blockers on nsPEF-induced membrane

depolarization. An analysis of variance indicated there was a significant difference between

the mean fluorescence between groups using different potassium blockers [F (6, 64) = 17.09,

p < 0.01, Ω2 = 0.60]. Post hoc analysis using the Games-Howell method revealed the

following: when comparing control conditions (nsPEF only) and those with potassium

channel blockers, only Penitrem A, a specific BK channel blocker [n = 5, X = 0.75, SE = 0.26,

p < 0.01, r = 0.96]; and 100 mM TEA, a broad voltage-gated potassium channel blocker (KV)

[n = 16, X = 0.58, SE = 0.26, p < 0.01, r = 0.96] significantly inhibited membrane

depolarization from a single 34 kV/cm, 10 nsPEF. Inhibiting ATP-dependent (KATP) channels

with Glibenclamide [n = 9, X = 2.73, SE = 0.27], small conductance voltage-and-calcium

dependent potassium (SK) channels with 5 mM TEA [n = 5, X = 2.42, SE = 0.18], intermediate

conductance voltage-and-calcium dependent potassium (IK) channels with 30 μM

Clotrimazole [n = 11, X = 2.55, SE = 0.09], or inward rectifying potassium channels (KIR) with

2.5 mM Barium chloride (BaCl2) [n = 10, X = 2.24, SE = 0.40] produced no significant

Page 73: Investigating the role of voltage-gated ion channels in

72

differences from controls. Asterisks * represents statistical significance at alpha = 0.05

compared to nsPEF only condition. Error bars represent +/- 2 SE

Whereas the 100 mM TEA condition significantly inhibited the nsPEF-induced

depolarization, the 5 mM TEA did not. Since it is known that KV inhibition at high

concentrations (> 10 mM) of TEA are known to include BK channels but not at low

concentrations (≤ 10 mM) [116], a range of TEA concentrations were used next to investigate

the dose-response of the TEA inhibition of the nsPEF depolarization. A statistically significant

trend can be seen, whereby increasing concentrations of TEA inhibited the nsPEF-induced

membrane depolarization (Fig. 3.5).

Figure 3.5 - Concentration-dependent inhibition of nsPEF depolarization by TEA.

Comparing the nsPEF-induced membrane depolarization measured using increasing

concentrations on tetraethyl ammonium (TEA) compared to the control conditions showed

significant group differences [F (5, 46) = 25.32, p < 0.01, Ω2 = 0.72]. Post hoc analysis using

Games-Howell test were conducted. Whereas 5 mM [n = 5, X = 2.42, SE = 0.18] TEA

concentrations did not have statistically significant effects on nsPEF-induced depolarization,

a linear trend can be seen as concentrations increase reaching statistical significance at 10 mM

[n = 7, X = 1.58, SE = 0.15, p = 0.02, r = 0.69], 20 mM [n = 5, X = 1.13, SE = 0.34, p = 0.02, r

= 0.87], 50 mM [n = 5, X = 0.68, SE = 0.22, p < 0.01, r = 0.84], and 100 mM [n = 15, X = 0.57,

Page 74: Investigating the role of voltage-gated ion channels in

73

SE = 0.08, p < 0.01, r = 0.96] TEA. Asterisks * represents statistical significance at alpha =

0.05 compared to nsPEF only condition. Error bars represent +/- 2 SE

Post-hoc analyses using Games-Howell reveal that statistically significant nsPEF

inhibition was observed in conditions using 10 mM, 20 mM, 50 mM, and 100 mM TEA

concentrations, but not with conditions using 5 mM TEA.

Since results from experiments using high concentrations of TEA and those using the

selective BK channel blocker, Penitrem A, suggest that BK channel blockers can inhibit the

nsPEF-induced membrane depolarization; two additional BK channel blockers were

investigated: 100 μM Paxilline and 10nM Iberiotoxin. When compared to the nsPEF only

condition BK channel block resulted in statistically significant inhibition of nsPEF-induced

depolarization (Fig. 3.6a). Post-hoc analysis using Games-Howell indicated that Paxilline but

not Iberiotoxin was able to significantly inhibit depolarization following nsPEF.

If the nsPEF application was depolarizing transmembrane potential via activating the

BK channel, as suggested by the inhibition of these effects by BK inhibitors, one would also

expect that activation of the BK channel by pharmacological agonists should produce similar

changes in PMPI as induced by nsPEF. To confirm this, a specific activator of the BK channel

was used without any exposure to pulsed electric fields. When 10 μM Phloretin was added to

gate the BK channel, a depolarization was observed that was statistically similar to that

observed following nsPEF (Fig. 3.6b).

Page 75: Investigating the role of voltage-gated ion channels in

74

Figure 3.6 - BK channel blockers significantly inhibit membrane depolarization

following 34 kV/cm nsPEF. a- An analysis of variance indicated there was a significant

difference in mean fluorescence between groups [F (4, 49) = 27.79, p < 0.01, Ω2 = 0.70]. Post

hoc analysis followed using the Games-Howell method. When comparing nsPEF only

conditions to those with Paxilline [n = 5, X = 0.35, SE = 0.25, p < 0.01, r = 0.94] and

Iberiotoxin [n = 15, X = 2.64, SE = 0.26], only the condition with 100 μM Paxilline significantly

inhibited the nsPEF-induced depolarization. b- A two-tailed t-test comparing relative

fluorescence measures indicated there was no significant difference between group means [t(2

tail)(11) = 0.32, p = 0.75] when 10 μM of Phloretin [n = 4, X = 3.26, SE = 0.14] was added to the

external media, compared to the response observed when cells were exposed to nsPEF [n = 9,

X = 3.10, SE = 0.32]. Data for tetraethyl ammonium chloride (TEA) and Penitrem A are

included from figure 4. Asterisks * represents statistical significance at alpha = 0.05 compared

to nsPEF only condition. Error bars represent +/- 2 SE

Page 76: Investigating the role of voltage-gated ion channels in

75

3.4.5 Depolarizing response of nsPEF is calcium-sensitive.

Among the potassium channel blockers explored, BK channel blockers were the only

ones able to significantly inhibit membrane depolarization following nsPEF application. Its

role was confirmed by using specific BK channel activators which mimicked the response

observed following pulse application. Given that the BK channel can be gated by voltage-and-

calcium, it was expected that modulating calcium should influence the sensitivity of the

response to nsPEF. To test whether modulating intracellular and extracellular calcium

concentrations could influence the threshold depolarization response following nsPEF,

experiments were conducted using the following parameters: calcium-free conditions were

performed with cells that were incubated with 10 μM BAPTA–AM and 1.8 mM Ethylene glycol

tetraacetic acid (EGTA) substituted for 1.8 mM calcium chloride; and conditions where 10 μM

of the calcium ionophore Ionomycin was added.

As represented in Fig. 3.7, altering the concentration of calcium shifted the threshold

of the electric field intensity required for depolarization. In the control condition (n = 25), the

EF50 (electric field required for 50 % maximal depolarization response) was determined to be

26.15 kV/cm where a maximal response was considered 100 %: A significant shift to the left

was observed in the Ionomycin condition (n = 19) where the EF50 was 20.34 kV/cm and the

maximal response exceeded the control by 10 %, whereas in the calcium-free condition (n =

21) the EF50 was shifted significantly to the right to a value of 45.43 kV/cm with the maximal

response being less than 50 % of that under control conditions. Descriptive statistics for these

results can be found in Table 3.3.

Page 77: Investigating the role of voltage-gated ion channels in

76

Figure 3.7 - Calcium-dependent threshold response curves in response to nsPEF

of varying electric field strengths. Removing extracellular and intracellular calcium

resulted in an EF50 (electric field intensity required to produce 50 % of the maximal

depolarization) of 45.43 kV/cm (blue trace) and reduced the maximal response to less than 50

% of that observed in control conditions where the EF50 was determined to be 26.15 kV/cm

(black trace). An opposite shift in EF50 of 20.34 kV/cm (red trace) was observed using the

calcium ionophore Ionomycin, which also resulted in a 10 % increase in maximal response

compared to control conditions. The overall fit for these dose-response curves was very good

with statistically significant differences between groups [χ2 (2) = 0.38, p < 0.01, Adj. R2 = 0.99].

Error bars represent +/- 2 SE

Page 78: Investigating the role of voltage-gated ion channels in

77

Electric field strength (kV/cm)

Control Calcium-free Ionomycin

Baseline n = 8, X = 1.87, SE = 0.47

n = 3, X = 0.93, SE = 0.40

n = 4, X = 2.58, SE = 0.76

16.5 n = 3, X = 2.47, SE = 1.02

n = 3, X = 1.56, SE = 0.21

n = 3, X = 10.38, SE = 1.70

22 n = 4, X = 30.73, SE = 10.78

n = 5, X = 5.42, SE = 1.22

n = 4, X = 75.30, SE = 19.38

34 n = 5, X = 91.93, SE = 8.56

n = 5, X = 25.40, SE = 3.06

n = 3, X = 100.94, SE = 14.83

44 n = 5, X = 100.00, SE = 8.56

n = 5, X = 46.01, SE = 10.60

n = 5, X = 114.03, SE = 9.42

Table 3.3 - Descriptive statistics from figure 3.7. Sample sizes (n), means (X), and

standard errors (SE) are included by calcium condition for each level of electric field strength

used.

Knowing that calcium was important for the nsPEF depolarizing effect, subsequent

experiments were undertaken to determine whether the source of the calcium; being either

extracellular (EC), intracellular (IC) or via release from the endoplasmic reticulum (ER)

played a role in the amplitude of the response. These experiments were separated into

extracellular calcium-free, intracellular calcium-free, and ER depleted; and these were

compared to both the controls and the calcium-free (IC+EC) results above. Cells were

incubated with 1.8 mM Ethylene glycol tetraacetic acid (EGTA), which was substituted for the

calcium chloride in the media for the EC calcium-free condition; whereas for the IC calcium-

free condition cells were incubated with 10 µM BAPTA-AM; and for the ER calcium-depleted

procedure cells were incubated with 1 µM Cyclopiazonic acid (Fig. 3.8).

Whereas conditions using 1 µM Cyclopiazonic acid were not significantly different than

controls, removing calcium from the extracellular and intracellular environment resulted in

significant inhibition that was statistically similar to those observed when removing EC and

IC calcium. Although removing EC and IC calcium resulted in significant inhibition of the

nsPEF-induced membrane depolarization, the greatest effect was observed when EC calcium

was removed.

Page 79: Investigating the role of voltage-gated ion channels in

78

Figure 3.8 - Comparing the role of calcium from intracellular, extracellular and

endoplasmic reticulum compartments. An analysis of variance indicated there was a

significant difference in average fluorescence between groups [F (4, 41) = 16.44, p < 0.01, Ω2

= 0.60]. Post hoc analysis followed using the Games-Howell method. Significant differences

in depolarization were observed when comparing controls with 1.8 mM calcium in media [n =

9, X = 3.10, SE = 0.32] to conditions where EC calcium was removed by substituting 1.8 mM

EGTA for calcium in the media [n = 7, X = 0.88, SE = 0.11, p < 0.01, r = 0.95], IC calcium was

removed by adding 10 µM BAPTA-AM [n = 10, X = 1.28, SE = 0.18, p < 0.01, r = 0.89], or both

[n = 5, X = 0.74, SE = 0.09, p < 0.01, r = 0.96]; however, no significant effect was observed

when calcium from the ER was depleted using 1 µM Cyclopiazonic acid [n = 11, X = 2.44, SE =

0.27]. Asterisks * represents statistical significance at alpha = 0.05 compared to nsPEF only

condition. Error bars represent +/- 2 SE

Page 80: Investigating the role of voltage-gated ion channels in

79

Next, to test whether the influence of calcium on the depolarizing response was solely

due to its role in modulating BK channel function or by nsPEF modulation of voltage-gated

calcium channels, a series of experiments using several calcium channel blockers were

performed (Fig. 3.9). Broad calcium entry channel blockers, Gadolinium, Lanthanum chloride

and Ruthenium Red were investigated; as were L-type calcium channel blocker Nifedipine and

T-type calcium channel blocker Mibefradil. Calcium can also enter via non-voltage-gated

calcium channels such as the family of transient receptor potential (TRP) channels; therefore,

selective TRPA1 blocker HC030031, TRPV1 blocker A784168, TRPC blocker 2-APB, and

TRPM8 blockers M8-B and AMTB hydrate were tested for the capacity to block the nsPEF

depolarization response.

Analysis revealed a statistically significant difference between conditions. Post hoc

analyses using Games-Howell method revealed a significant inhibition of membrane

depolarization between 10 µM Nifedipine, 20 µM Mibefradil, combined 10 µM Nifedipine and

20 µM Mibefradil, 100 µM Gadolinium, 100 µM Ruthenium Red, 100 µM Lanthanum

chloride, 100 µM 2-APB, 50 µM M8-B, and 50 µM AMTB hydrate conditions when compared

with controls. Conversely, no significant differences were observed between control conditions

and those using 300 µM HC030031 or 10 µM A784168.

Page 81: Investigating the role of voltage-gated ion channels in

80

Figure 3.9 - Effects of calcium channel blockers on nsPEF-induced membrane

depolarization. An analysis of variance revealed a significant difference in mean

fluorescence between groups [F (11, 91) = 11.71, p < 0.01, Ω2 = 0.56]. Post hoc analysis were

then performed using the Games-Howell method. Conditions using 10 µM of TRPV1 blocker

A784168 [n = 9, X = 3.43, SE = 0.42], or 300 µM of TRPA1 blocker HC030031 [n = 6, X =

2.38, SE = 0.41] produced no significant differences from control conditions [n = 9, X = 3.10,

SE = 0.32]. Significant inhibition of nsPEF-induced membrane depolarization was observed

in conditions using broad calcium channel blockers 100 µM Gadolinium [n = 9, X = 1.69, SE

= 0.12, p = 0.04, r = 0.70], 100 µM Ruthenium Red [n = 7, X = 1.43, SE = 0.23, p = 0.04, r =

0.71] and 100 µM Lanthanum chloride [n = 7, X = 1.07, SE = 0.33, p = 0.02, r = 0.83]; with

the selective L-type calcium channel blocker 10 µM Nifedipine [n = 6, X = 1.41, SE = 0.14, p =

0.02, r = 0.90], T-type calcium channel blocker 20 µM Mibefradil [n = 8, X = 0.81, SE = 0.23,

p < 0.01, r = 0.91], or when 10 µM Nifedipine was combined with 20 µM Mibefradil [n = 4, X

= 0.71, SE = 0.29, p < 0.01, r = 0.93]; as well as with TRPC blocker 100 µM 2-APB [n = 9, X =

1.57, SE = 0.19, p = 0.03, r = 0.69], or TRPM8 blockers 50 µM M8-B [n = 5, X = 0.88, SE =

0.28, p < 0.01, r = 0.91] and 50 µM AMTB hydrate [n = 9, X = 0.44, SE = 0.18, p < 0.01, r =

0.95]]. Asterisks * represents statistical significance at alpha = 0.05 compared to nsPEF only

condition. Error bars represent +/- 2 SE

Page 82: Investigating the role of voltage-gated ion channels in

81

3.4.6 nsPEF-induced membrane depolarization is not mediated by voltage-gated Na+ channels; however, Na+ ions may still be involved.

With evidence that nsPEF depolarization responses were abrogated by inhibitors of

voltage-gated K+ and voltage-gated Ca2+ channels, the following section investigated inhibitors

of voltage-gated Na+ and Cl- channels. For these experiments 1 µM of Tetrodotoxin (TTX) and

200 µM of Amiloride hydrochloride were used to block voltage-dependent Na+ channels.

Further tests were performed by removing extracellular Na+ and replacing it with equimolar

140 mM N – methyl – D - glucamine (NMDG) or Choline chloride.

The results from this series of investigations revealed (Fig. 3.10) a significant difference

in group means. It should be noted that the assumption of homogeneity of variance was not

met as determined using Levene’s test; therefore, Welch’s F test was used as a correction. Post

hoc analysis using Games-Howell test showed no significant differences between conditions

using TTX or Amiloride HCl when compared with controls. In contrast, removing extracellular

sodium using either NMDG or Choline chloride significantly inhibited membrane

depolarization caused by nsPEF.

These results, however, are difficult to interpret given that both NMDG and Choline

chloride are known to block BK channel activity [122, 123]. Confirmation of this fact is evident

in the same analysis where 10 µM Phloretin was applied in the presence of Choline chloride

and the depolarizing effect observed above, which was on the same order of magnitude as that

observed following nsPEF delivery, was abolished.

Page 83: Investigating the role of voltage-gated ion channels in

82

Figure 3.10 - Comparing effects from sodium and chloride channel blockers on

nsPEF-induced membrane depolarization. An analysis of variance, reporting Welch’s

F due to a violation of the assumption of homogeneity of variance, revealed a significant

difference in mean depolarization responses between groups [F (6, 20.3) = 61.69, p < 0.01, Ω2

= 0.70]. Post hoc analysis using the Games-Howell method found no significant effects on

depolarization response following nsPEF delivery when comparing control conditions [n = 9,

X = 3.10, SE = 0.32] to those with voltage-gated sodium channel blockers 1 µM TTX [n = 7, X

= 2.96, SE = 0.25] and 200 µM Amiloride hydrochloride (HCl) [n = 8, X = 2.93, SE = 0.69].

Conversely, when sodium was replaced in the extracellular media there was significant

inhibition of the membrane depolarization using either N – methyl – D - glucamine (NMDG)

[n = 9, X = 0.53, SE = 0.11, p < 0.01, r = 0.96] or Choline chloride [n = 9, X = 0.51, SE = 0.05,

p < 0.01, r = 0.97]. Repeating the experiments with Choline chloride in the presence of the BK

channel activator Phloretin 10 µM resulted in inhibition of the depolarizing response [n = 10,

X = 0.04, SE = 0.02, p < 0.01, r = 0.98]. Error bars represent +/- 2 SE

Page 84: Investigating the role of voltage-gated ion channels in

83

3.4.7 nsPEF-block is reversible.

In order to demonstrate that the inhibitors used were in fact inhibiting the nsPEF-

induced membrane depolarization, we first had to rule out the possibility that the inhibitors

were depolarizing the cells during the incubation period. If they were indeed depolarizing the

cells they would not have the ability to further depolarize in response to nsPEF, which could

explain nsPEF depolarization-inhibition. This possibility was ruled by applying multiple

pulses at five minute intervals and observing the response over time. After three initial pulses,

the imaging solution containing the inhibitor was washed off and replaced with the control

solution (Fig. 3.11).

Figure 3.11 - Reversibility of the TEA and Penitrem A inhibition of the nsPEF-

induced membrane depolarization. A single 34 kV/cm nsPEF (n = 9) was delivered every

5 minutes (indicated by black arrows in figure) while the TEA (n = 15)/Penitrem A (n = 6)

(solution was replaced with control solution after 17.5 minutes (indicated by black bar in

figure) demonstrating that the inhibition of the nsPEF-induced depolarizing response was not

due to any depolarizing effect of inhibitors during incubation. For the purpose of clarity, only

the 10 µM Penitrem A and 100 mM TEA data is displayed. Reversibility of the response was

consistent among other inhibitors. Error bars represent +/- 2 SE

Page 85: Investigating the role of voltage-gated ion channels in

84

Following the removal of the pharmacological inhibitor, subsequent nsPEF delivery

produce depolarizing response expected in the control conditions. This reversibility of the

nsPEF block demonstrates that the cells were not depolarized during incubation and that the

inhibitors used were in fact inhibiting the plasma membrane depolarization following nsPEF

application. Another important feature is that repolarization of the resting membrane

potential could be prevented by applying a subsequent pulse every five minutes.

3.5 Discussion

The purpose of this investigation was to study the role of voltage-gated ion channels in

the depolarization response of the transmembrane potential following delivery of a single 10

nanosecond pulsed electric field. A combination of pharmacological inhibitors and a

fluorescent indicator of plasma membrane potential was used with live cell imaging to evaluate

the role of the various ion channels expressed in human U87 glioblastoma cells. These intense,

short duration pulsed electric fields are known to exert their influence on cell physiology by

inducing a transmembrane voltage across the cell leading to the formation of pores in the

plasma membrane and intracellular organelles. This study implicated several ion channels in

the nsPEF response and pharmacologically discriminated the specific voltage-gated channels

involved.

The membrane potential indicator (PMPI) we employed in our investigations has

previously been used in several studies and has been shown to be a valuable tool for

monitoring membrane potential with higher throughput than patch clamp electrophysiology

[77, 124, 125]. Our calibration of PMPI fluorescence with transmembrane potential was

consistent with previous studies that compared PMPI fluorescence using electrophysiology

[126].

By using a wide range of ion channel blockers and ion substitutions, we have provided

strong evidence that the application of a single nsPEF is directly modulating ion channels

within the parameters used in this study (10 ns pulse, 34 kV/cm). With that said, these results

Page 86: Investigating the role of voltage-gated ion channels in

85

may at first appear to be somewhat contrary to what one would expect. It is counterintuitive

that the BK channel would gate at a negative resting membrane potential in the absence of

Ca2+ as we have observed with nsPEF depolarization; however, this has been reported in non-

excitable cells and the mechanism is well characterized [127]. The BK channel is primarily a

hyperpolarizing channel and blocking the BK channel would be expected to increase the

depolarizing effect observed; however, multiple studies have demonstrated that blocking the

BK channels could have an inhibitory effect on depolarization and neurotransmitter release

[128–130]. Interestingly, the BK channel has been demonstrated in some cells to operate in a

biphasic mode [131, 132]. In excitable cells such as neurons, the activation of BK channels

induces a hyperpolarizing response leading to a decrease in intracellular calcium and closing

of voltage-gated calcium channels creating a negative feedback loop. Conversely, in non-

excitable cells activation of BK results in increased intracellular calcium through activation of

non-voltage gated calcium channels thereby increasing the driving force of calcium, and

creating a positive feedback loop.

It is worth noting that the dimensions of punctate depolarized clusters and “hotspots”

observed in PMPI labelling are consistent with previously measured size and distribution of

lipid rafts within the cell membrane [133, 134]. From a physiological perspective, it is well

established that many ion channels including voltage-gated ion channels are commonly

clustered in lipid rafts [117, 134]. Specifically the BK channel has been shown to form

complexes with T-type calcium channels [135] and L-type calcium channels [136].

The positive-feedback relationship between calcium channels and BK channels

explains the results presented in this work and can be summarized as follows: Activation of T-

type and L-type voltage-gated calcium channels by the nsPEF generates an inward calcium

current which results in activation of BK channels. Activation of voltage-gated calcium

channels has been confirmed previously with 5 ns, 50 kV/cm pulses [64] and shorter 500 ps,

190 kV/cm pulses [63]. Here, BK activation creates an outward potassium current which can

activate calcium influx through non-voltage-gated calcium channels [131, 132], in the case of

Page 87: Investigating the role of voltage-gated ion channels in

86

U87 cell, where TRPM8 channels are known to interact with BK channels in glioblastoma cells

[137]. It is also possible that the nsPEF could simultaneously activate BK channels and voltage-

gated calcium channels. The fact that using a specific BK channel activator could depolarize

the cells to a similar degree as the nsPEF corroborates this possibility (Fig. 3.12).

Figure 3.12 - Proposed mechanism of direct interaction between nsPEF and

voltage-gated channels along with downstream effects on non-voltage dependent

channels. 1. A single 10 ns, 34 kV/cm nsPEF activated voltage-gated calcium channels (L and

T-type) along with the voltage- and calcium-dependent BK channels, causing an influx of Ca2+

and efflux of K+. 2. The K+ efflux then triggers further Ca2+ entry through non-voltage-gated

cation channels (TRPM8). 3. This action establishes a positive feedback loop and results in

further membrane depolarization.

Considered the “gold standard” for studying BK channel activity, the lack of inhibitory

action of the selective BK channel blocker, Iberiotoxin, was an unexpected result. However,

reviewing the literature shed light on this apparent mystery. Toxin resistant BK channels have

Page 88: Investigating the role of voltage-gated ion channels in

87

been described since the late 1980’s [138]. Originally found to be insensitive to Charybdotoxin,

these channels are also insensitive to Iberiotoxin, which is a closely related toxin [139].

As a general overview of the BK channel, it consists of a tetrameric assembly of alpha

subunits (BKα), which are responsible for creating the pore forming domain, and also contain

the calcium biding domain and voltage sensing domain. Co-expressed with the alpha subunits

are auxiliary beta subunits which are known to modulate the channel’s activity [132, 140, 141].

The resistance to toxins is due to the presence of a specific isoform of the beta subunit (β4),

which is known to be widely expressed in the brain [142–144]. Moreover, glioblastoma cells

are known to express a specific subtype of BK channel known as the glioma BK or gBK [117],

which incidentally are also known to express the Iberiotoxin-and-Charybdotoxin resistant β4

subunit [145].

It was not possible to definitively rule out Na+ in the nsPEF transmembrane

depolarization response, although it was clear from our results that voltage-gated Na+

channels were not involved. When Na+ was removed from the extracellular media and replaced

with equimolar NMDG+ or Choline+, the depolarizing effect of Phloretin was eliminated. This

result makes sense as both NMDG+ and Choline+ are also known to inhibit BK channel activity

[122, 123, 146, 147]. It is of interest that TRPM8 channels were found to have a role in the

nsPEF depolarization response as they are permeable to divalent and monovalent cations

[148–151]. BK activation with TRPM8 gating would undoubtedly transport Na+ in addition to

Ca2+ into the cell and this would be commensurate with recent electrophysiological studies

that have identified Na+ currents in the early stage of nsPEF effects [152].

The involvement of TRP channels in the depolarization response caused by nsPEFs

was also an important finding as some of the TRP channels are known to dilate to form pores

permeable to small cationic fluorescent molecules such as YO-PRO-1 and ethidium. TRPA1,

TRPV1 and TRPM8 are cell sensors for thermal, chemical and mechanical stimuli that are also

known to form large, stable and reversible pores [149, 153], intriguingly similar to those caused

by nsPEFs. Our investigation showed TRPM8 inhibition partially abrogated the nsPEF

Page 89: Investigating the role of voltage-gated ion channels in

88

depolarization response. It is still controversial whether TRPM8 form large pore complexes

[154], and pore dilation is more confirmed with TRPA1 and TRPV1 [154–156]; so future studies

should consider the influence of nsPEFs on pore-dilating TRPs as these are also known to be

upregulated in cancer [113, 137].

BK channels (Big Potassium, also called Maxi-K or slo1), are a potentially important

target for nsPEF effects as they have currents in order of 100-300 picoSiemens [157] and they

are known to be upregulated in glioma and many other types of cancer [158, 159]. BK

activators are also known to lead to apoptosis [160]. Mechanistically, it is not clear whether

electric pulses in the nanosecond timescale could directly open BK channels as their gating

time at physiological voltages requires multiple states with an activation time constant in the

order or 150-200 microseconds [161]. Certain isoforms of BK channels are also known to be

sensitive to mechanical stimulation and membrane stretch [162]. Further investigations

should consider the mechanical properties of both BK and TRP channels in nsPEF effects,

given that these stimuli have recently been shown to generate pressure transients [163].

Finally, we have also demonstrated that a single nsPEF delivered at five minute

intervals can depolarize and prevent reestablishment of the resting membrane potential of

U87 glioblastoma cells. Experiments are underway exploring the long-term manipulation of

transmembrane potential to determine if non-electroporating nsPEFs can disrupt cell

physiology in a manner that may be therapeutically useful for the treatment of cancer. The

repolarization of plasma membrane potential in cancer cells repeatedly depolarized by nsPEF

would be expected to occur via the Na+/K+-ATPase at great energetic cost, and given that only

cancer cells expressing voltage-gated ion channels would be sensitive to this treatment, this

may represent a new means to bioelectrically exhaust cancer, while sparing non-malignant

cells nearby.

3.6. Conclusion

The primary aim of this section was to investigate the possibility that nsPEF are

acting on transmembrane ion channels. A single 10 nsPEF was used at varying electric field

Page 90: Investigating the role of voltage-gated ion channels in

89

intensities to explore this. We monitored the transmembrane potential using the commercial

indicator, PMPI, which proved to be a useful tool for studying long-term effects. Using a

variety of pharmacological modulators of ion channel activity, we found that voltage-gated

potassium and calcium channels were intimately connected to the observed depolarization

following nsPEF exposure, as was the TRPM8 cation channel. Blocking these channels

resulted in a significant decrease in membrane depolarization post-nsPEF exposure. We

provided a potential theoretical model explaining how the interconnectivity between these

channels, that are often found in proximity in lipid rafts, could help explain our results.

Because the expression of ion channels varies significantly among cell lines, nsPEF effects

would most likely be explained by different ion channels in different cells.

Page 91: Investigating the role of voltage-gated ion channels in

90

Chapter 4 -

Plasma membrane depolarization

and permeabilization due to electric

pulses in cell lines of different

excitability

Page 92: Investigating the role of voltage-gated ion channels in

91

4.1 – Introductory remarks

The following chapter has been adapted from a manuscript that is currently under

revision. This was a collaborative project between our lab and the University of Ljubljana.

Results presented will include only those to which I directly contributed. Three cell lines were

used for the following experiments: U87 glioblastoma cells, CHO cells, and HT22 mouse

hippocampal neurons. HT22 cells were used in their differentiated and non-differentiated

states. Experiments were conducted using pulse durations ranging from 10 ns up to 10 ms,

where membrane potential and membrane integrity were the outcome variables measured.

For both variables, the goal was to determine the threshold, or the lowest electric field

intensity, required to depolarize or permeabilize the cells.

The bigger picture for these experiments was to test the potential for optimizing

current treatments using electropermeabilization, such as electrochemotherapy. One common

complaint for electrically based therapies is the associated discomfort due to the electrical

excitation of the surrounding muscle and nervous tissues, which leads to muscle contractions

and pain. We investigated whether it would be possible to find an electric field intensity that

could simultaneously permeabilize the target tissues while minimizing the excitation of the

surrounding nerves and muscles.

We found a statistically significant difference in excitability between cell types.

Although the electric field threshold required to permeabilize the cells did not differ

significantly between cell lines, the differentiated neurons required a stronger electric field to

depolarize their transmembrane potential compared to the other cell lines. Although this was

an in vitro study, these results indicate that it may be possible to modulate the electric field

intensity in a clinical setting to enhance permeabilization, while limiting the excitation of

neighboring nerves and muscles, thus the sensation of pain.

Page 93: Investigating the role of voltage-gated ion channels in

92

4.2 Introduction

Short, high-voltage pulses have been shown to increase the permeability of cell membranes

to different molecules (reversible electroporation) or cause cell death (irreversible

electroporation - IRE) [164–166]. Electropermeabilization is used in biotechnology sectors for

food-processing [167–169], and in medicine [170] for gene electrotransfer (GET) [37, 48, 171],

DNA vaccination [172–175], transdermal drug delivery [176, 177], IRE as a soft tissue ablation

technique [42, 45, 46, 178] and electrochemotherapy (ECT) [179–182].

Medical applications of electroporation extend to a variety of tissues and tumours. Some

of these tissues are electrically excitable, such as neurons in the central and peripheral nervous

systems as well as muscle tissues. These cells are known to undergo sponataneous electrical

activity [183] which is how they communicate with other tissues, and these excitable cells may

be particularily vulnerable to electrically-based treatments.

In the literature, there are several examples of ECT, IRE and GET targeting excitable

tissues using electric pulses. Brain cancer has been treated with IRE, and electric pulses have

been shown to transiently disturb the blood-brain-barrier allowing chemotherapeutics to

enter the brain [44, 184–188]. Treating prostate cancer [43, 189], bone metastases [179, 190],

and tumours in the spine can affect the surrounding nervous tissue [190]. When treating

tumours in other parts of the body, electrodes will invariably be in the vicinity of the nerves or

muscles where the electric field is high enough for excitation or even permeabilization. Electric

pulses are also used for ablation of myocardial tissue to treat atrial fibrillation [191–193].

Muscles are a popular target for gene electrotransfer as they are easily accessible and

transfected [194, 195]. Among them, the heart can be electroporated to treat ischemia [196,

197].

Several studies have shown that the effect of electric pulses on the functionality of excitable

tissues was only short-term. Following IRE, nerves in different animal models recovered

electro-physiologically, histologically and functionally [190, 198–200] or at least showed a

Page 94: Investigating the role of voltage-gated ion channels in

93

potential for regeneration [201]. After electroporation of neurons within the rat neocortex , in

vitro and in vivo, the membrane potential, the action potential waveform and passive

membrane properties remained unchanged [202]. Following pulmonary vein ablation using

electroporation, the histology and functionality of phrenic nerve remained unchanged [203].

Finally, no histological damage on nerves in the neurovascular bundle was observed post-IRE

treatment of the prostate [199].

Some of the main drawbacks to treating tissues with pulsed electric fields are; the pain

associated with repeated electrical stimulation [204–207], the need to administer muscle

relaxants and anaesthesia [208] and synchronization with the ECG [209–211]. The neurons

responsible for pain sensation are also known as nociceptors and have been shown to be

stimulated by electric pulses [212, 213]. An important advancement for PEF treatments would

be to determine a point at which maximum permeability of the membrane could be achieved

while minimizing excitation of the nearby excitable cells. In ECT, for example, this would

translate to maximum drug delivery into tumour cells with minimum tissue damage to

surrounding regions, reduced pain experienced by the patient, and minimal use of muscle

relaxants.

The following experiments will explore how pulsed electric fields will effect excitable and

non-excitable cells. This will be addressed using measures of membrane potential and

membrane permeability for pulse durations ranging from 10 ns to 10 ms. Using a range of

electric field intensities, the goal will be to determine the lowest intensity required to

depolarize the transmembrane potential and to permeabilize the plasma membrane. The

results will be compared between cell types to determine whether or not it could be possible

to optimize current treatments to reduce the sensation of pain due to repeated electrical

stimulation.

Page 95: Investigating the role of voltage-gated ion channels in

94

4.3 Materials and Methods

Some of the materials and methods are common between thesis chapters. These include

cell culture, growth and maintenance; fluorescence microscopy, pulse generators and

electrodes used; data extraction from images, as well as statistical methods used to analyze

data. In the interest of not repeating information, these have been explained in detail in the

general methods section of the introductory chapter.

4.3.1 Cell culture and preparation

Three cell lines were used (Fig. 4.1) for the following experiments. CHO Chinese

hamster ovary cells (Fig. 4.1a), U87 human glioblastoma cells (Fig. 4.1b), and HT22

immortalized mouse hippocampal neurons. The HT22 cell line was used in their non-

differentiated and differentiated states. Following differentiation, morphological changes

were evident and the cells stopped dividing (Fig. 4.1 c vs d).

4.3.2 Potassium calibration of PMPI.

A chemical calibration was performed for all cell types to determine their degree of

excitability. This was accomplished by measuring membrane depolarization (PMPI

fluorescence) in conditions with increasing [K+] in the imaging solution (2.5 mM, 25 mM, 50

mM, 75 mM, 100 mM, 140 mM). Osmolarity was maintained by removing NaCl in an

equimolar manner as KCl was increased. Each sample was monitored continuously with the

buffer being replaced every 5 minutes.

Page 96: Investigating the role of voltage-gated ion channels in

95

Figure 4.1 - Phase-contrast images of all four cell lines used in experiments. a –

CHO; b - U87 MG; c - undifferentiated HT22; and d - differentiated HT22. All images were

taken at 200x magnification.

4.4 Results

4.4.1 Cell Excitability

The first experiment compared how each cell line responded to chemical depolarization by

increasing the extracellular [K+] over time (Fig. 4.2). Here we found that the greatest relative

change in membrane potential occurred in the CHO cell line. U87 cells and the

undifferentiated HT22 cells responded almost identically, whereas the differentiated HT22

cells showed the least amount of relative change. The four cell lines began to show the greatest

divergence when the 50 mM [K+] buffer was added.

Page 97: Investigating the role of voltage-gated ion channels in

96

Figure 4.2 – Chemical depolarization of cells using K+. Fluorescence changes were

monitored in all four cell lines. The imaging solution was changed every 5 minutes with

increasing [K+] ranging from 2.5 mM in the original imaging solution up to 140 mM. Error

bars represent +/- 2SE.

A representative response to the depolarization dynamics following pulsed electric field

exposure can be seen in (Fig. 4.3). A similar trend was observed for all pulse durations, where

the U87 cells showed the greatest depolarization and the differentiated HT22 cells showed the

least. One exception to this was in the 10 ms condition, where the CHO cells were depolarized

more than all other cells. The maximal depolarization was observed within the first 2.5

minutes following pulse exposure, after which the membrane potential gradually returned to

near baseline levels. The time for repolarization differed based on the magnitude of

depolarization but ranged anywhere from 5 – 25 minutes.

Page 98: Investigating the role of voltage-gated ion channels in

97

Figure 4.3 – Representative depolarization dynamics following PEF

exposure. Membrane depolarization was observed after a single 10 ns PEF of 44 kV/cm,

with the exception of the differentiated HT22 cells requiring a minimum of 52 kV/cm to

depolarize the transmembrane potential. Error bars represent +/- 2SE.

Figure 4.4 – Magnitude of depolarizing response to pulsed fields from 10 ns

– 10 ms. U87 cells show a greater response to PEF-induced membrane depolarization

compared to all cells, but they also show the greatest variability. Error bars represent +/-

2SE.

Page 99: Investigating the role of voltage-gated ion channels in

98

Another interesting feature is highlighted in Fig. 4.4, where the U87 cells not only show

the largest depolarization response, but they also show the greatest variability in their

response. In some cases, such as the 1 µs duration, U87 cells had 5 times more variability than

the other cell lines. An ANOVA was performed comparing the variance between cells, which

confirmed that the variability in the U87 cell line was statistically greater than the other cells

[F(3, 27) = 4.19, p = 0.01, Ω2 = 0.24].

Fig. 4.5 shows the strength-duration curve for depolarization of all cell lines. The exact

values are listed in Table 4.1. We can see that an inverse relationship between the threshold

and pulse duration is evident, such that longer pulses require a much lower electric field to

elicit a depolarizing response. The statistical parameters for the analyses of the depolarization

threshold are shown in the Supplementary data section (Table S1 and S2).

Table 4.1 - The depolarization thresholds for all tested pulse durations and cell

lines. The results listed here are a tabular representation of the data in Fig. 4.5. The asterisk

(*) denotes that the threshold for this condition had to be estimated due to the large variability

of the data. All data is presented in kV/cm.

Electric field (kV/cm)

Pulse duration 10 ns 550 ns 1 μs 10 μs 100 μs 1 ms 10 ms

CHO 44 2.0 1.2 0.45 0.30 0.15 0.10

U-87 MG 34 2.2 1.4 0.60 0.35 0.20 0.12

Undifferentiated HT22

34 2.0 1.4 0.70 0.50 0.35 0.24

Differentiated HT22

52 2.0 1.7* 0.90 0.60 0.40 0.28

Page 100: Investigating the role of voltage-gated ion channels in

99

Figure 4.5 - The strength-duration curve for depolarization thresholds of all cell

lines. The minimum electric field intensity required to depolarize the transmembrane

potential is given for all pulse durations tested in kV/cm. As the pulse duration increases, the

electric field required to depolarize the cell decreases.

4.4.2 Plasma Membrane Permeability

The normalized plasma membrane permeabilization curve to YO-PRO is shown for all four

cell lines in Fig. 5.6. The threshold for electropermeabilization (Fig. 4.6a) in U87 and CHO

cells was 0.4 kV/cm, whereas it was slightly higher (0.6 kV/cm) for the undifferentiated and

differentiated HT22 cells. A similar increase in permeabilization was seen for all cell lines as

the electric field strength increased. The permeabilization curve could be best described using

a symmetric sigmoid (Fig. 4.6b). Table 4.2 summarizes some of the parameters from the curve,

which show that all four cell lines reached 50% permeabilization (E50%) between 0.8 and 0.9

kV/cm.

Page 101: Investigating the role of voltage-gated ion channels in

100

Figure 4.6 - Normalized permeabilization curve of all four cell lines to YO-PRO,

5 min after the pulse application. The threshold for electroporation was reached at

0.4 kV/cm (U-87 MG and CHO) or 0.6 kV/cm (undifferentiated and differentiated HT22

cells). a) Normalized values, the error bars represent one standard deviation and b) the fitted

symmetric sigmoid. Normalized fluorescence fn/- is presented as a function of applied electric

field E/(kV/cm).

An example of the time dynamics for YO-PRO uptake following 8 x 100 μs pulses at

1.2 kV/cm is shown in Fig. 4.7a. The maximal fluorescence and the resealing dynamics (τ) were

extracted by fitting a first-order uptake model. The maximal fluorescence (Fig. 4.7b) was the

highest for the U-87 MG cells and the lowest for the CHO cells. There were no significant

differences in maximal fluorescence between the differentiated and undifferentiated HT22s,

while all other pairwise comparisons using a t-test yielded statistically significant differences.

The value of time constant τ (Fig. 4.7c) corresponds to the time when 63% of the pores in the

membrane resealed. The resealing was the fastest for the U-87 MG cells, similar for the

undifferentiated HT22s and CHO cells, and slowest for the differentiated HT22 cells.

Page 102: Investigating the role of voltage-gated ion channels in

101

Figure 4.7 - Time dynamics of YO-PRO uptake and analyses. a - Time dynamics of

the YP uptake for all four cell lines following 8 x 100 µs pulses at 1.2 kV/cm and 1 Hz repetition

frequency. Error bars represent one standard deviation. Fluorescence (arbitrary units) is

presented as a function of time (min). b - The maximal value of YP fluorescence (arbitrary

units) for all four cell lines. c - The resealing constant (τ/s) is presented for each cell line. Error

bars for b and c represent the 95% confidence interval.

Cell line E50% (kV/cm) b (kV/cm) R-squared

CHO 0.80 ± 0.06 0.15 ± 0.11 0.98

U-87 MG 0.81 ± 0.10 0.13 ± 0.09 0.99

HT22 undifferentiated 0.94 ± 0.07 0.12 ± 0.06 0.99

HT22 differentiated 0.91 ± 0.07 0.12 ± 0.06 0.99

Table 4.2 - Parameters of the fitted symmetric sigmoid to the normalized data of

YO-PRO uptake. The numbers listed denote the optimal value and the corresponding 95%

confidence intervals.

4.5 Discussion

The goal of this study aimed to compare the depolarization thresholds between excitable

and non-excitable cells. For each cell, the strength-duration curve was determined, using the

obtained threshold values, following exposure to a single pulse ranging in duration from 10 ns

Page 103: Investigating the role of voltage-gated ion channels in

102

to 10 ms. The permeability curve was also determined for each cell line following application

of 8 x 100 µs pulses, delivered at 1 Hz. For the assessment of cell depolarization we used the

PMPI dye. For the assessment of the plasma membrane permeabilization we used YO-PRO

(YP) dye.

PMPI is a valuable tool for the measurement of membrane potential [125] and ion channel

pharmacology [214, 215]. Although electrophysiology is considered the gold standard for

measurement of membrane potential, PMPI has several advantages, namely the ease of use,

the ability to monitor long-term changes, and the ability to monitor multiple cells

simultaneously. Furthermore, PMPI has been compared directly to electrophysiology data

with a good agreement [126]. It consists of a two-part system which includes a fluorescent

anionic voltage-sensor and a quencher. When a cell is depolarized, the sensor translocates

across the plasma membrane resulting in an increase in fluorescence. Conversly, the

quenching molecule is excluded from the cell. With that said, when the plasma membrane is

permeabilized, it could be possible for the quenching molecule to enter the cell through pores

formed in the membrane and decrease the fluorescence in the cell. Interestingly, when pulses

were applied well above the depolarization threshold, we observed a decrease in the

fluorescent signal which could be indicative of cell electroporation. In the future, it should be

established to what extent the PMPI dye could also serve as cell membrane permeabilization

indicator.

The results following electrochemical depolarization show that with increasing K+

concentration, the fluorescence and thus the transmembrane voltage are increasing which is

in agreement with theory. It was unexpected, however, that the highest fluorescence was

achieved in CHO cells and not with the differentiated HT22 cells. Excitable cells typically have

a higher density of voltage-gated channels, thus more ions should enter the cell when these

channels are open. CHO cells are non-excitable cells and would be expected to have a low

expression of voltage-gated ion channels; however, some reports indicate that these cells

express voltage-gated Na+ channels [216] and voltage-gated Ca2+ channels [217].

Page 104: Investigating the role of voltage-gated ion channels in

103

CHO cells have a resting membrane potential around - 10 mV [218], which is similar to

that of U87 cells [116]. Although electrophysiology studies on HT22 cells were not found, a

typical neuron has a resting potential around – 70 mV [219]. If we recall from the previous

chapter, the depolarizing response measured here would be associated with a change in

magnitude less than 30 mV. One possibility could be that the resting potential of the neurons

did not reach the threshold required to activate voltage-gated channels.

The relatively low resting potential of the U87 cells combined with the large complement

of voltage-gated channels [116] could explain why the U87 cells experienced the greatest

depolarizing effect to one pulse. It is also worth noting that cancerous tissues have been shown

to be more affected by PEFs than non-cancerous tissues [220, 221].

The repolarization time for all four cell lines was in the range of minutes. Certainly this

was most surprising when considering the differentiated neurons which normally would be

expected to have repolarization times in the millisecond range [219]. There are several possible

explanations for the longer-than-expected repolarization time. First, as the assessment

method, we used PMPI dye, which has a time constant of several seconds [222]. Second, our

experiments were performed at room temperature which would be expected to slow down the

kinetics of the Na+/K+ pump. Third, it is possible that cells were depolarized as well as

electroporated. If true, electroporation has been reported to cause leakage of ATP [223] which

is necessary for repolarizing the membrane. Fourth, due to such a high induced

transmembrane voltage (several volts), it is possible that ion channels could be damaged

[224].

The time required for reaching the peak fluorescence in depolarization experiments

coincided with the resealing time observed in the permeabilization experiments. It is possible

that during depolarization and permeabilization experiments, PMPI and YP were entering

through voltage-gated channels [154] as well as through pores formed in the plasma

Page 105: Investigating the role of voltage-gated ion channels in

104

membrane. Even when using channel inhibitors, a total inhibition of depolarization could not

be achieved which indicates that during depolarization ions also enter through pores [100].

In the next part of our study, we exposed cells to 8 x 100 μs pulses, which are typically used

in electrochemotherapy treatments. All four cell lines reached the threshold of electroporation

at approximately the same value - between 0.4 and 0.6 kV/cm. The permeabilization curve of

all four lines could be described using a symmetric sigmoid. Although differentiation of HT22

cells causes a drop in the resting membrane potential [225], the more negative resting

membrane potential did not affect the threshold for electroporation. The permeabilization

curves followed similar dependency and reached 50% of the maximal fluorescence around

0.9 kV/cm. We can conclude that irrespective of the cell’s excitability, all four cell lines

responded similarly to electroporation pulses.

The YO-PRO uptake was the greatest in the U87 cell line and least in CHO cells. As

mentioned previously, cancerous cell have been shown to be more susceptible to PEFs when

compared to non-cancerous cells which could explain why they were the most permeabilized.

Since the cells were grown attached in a monolayer, the reason CHO cells showed the least

YO-PRO uptake could be due to the tendency of CHO cells to grow in colonies in close

proximity which decreases the area of the plasma membrane available for dye uptake. The

close proximity of cells could have also decreased the induced transmembrane potential due

to shielding [226–228].

4.6 Conclusion In summary, the depolarization threshold was higher for the excitable cells than for the

non-excitable cells. All four cell lines responded similarly to pulses of standard

electrochemotherapy parameters. The shape of the permeability curve was similar to curves

already published in the literature [229]. Thus, electroporation is a feasible means of treating

excitable and non-excitable cells with pulses of similar parameters. Furthermore, our results

show the potential of achieving permeabilization and minimizing or avoiding excitation/pain

Page 106: Investigating the role of voltage-gated ion channels in

105

sensation which needs to be explored in more detail. In future studies, it should be established,

however, to what extent in vitro depolarization and excitability correlate to the actual

excitation and pain sensation in vivo.

Page 107: Investigating the role of voltage-gated ion channels in

106

Supplementary Data

Table S1 - Statistical parameters for the strength-duration curve by cell line. F(x,

y) = ANOVA score with degrees of freedom in parentheses. Fw = Welch’s F adjustment when

the assumtion of homogeneity of variance was not met. The statistical significance provides

the p-values which are compared to our alpha criterion of α = 0.05. The effect size column

provides the proportion of variation in response attributed to the PEF.

CHO

ANOVA Significance Effect size (Ω2) 10 ns F (2 , 17) = 9.00 p < 0.01 0.47 550 ns F (3 , 15) = 4.56 p = 0.02 0.36 1 µs F (3 , 16) = 91.52 p < 0.01 0.93 10 µs FW (3 , 6.08) = 14.12 p < 0.01 0.67 100 µs F (4 , 19) = 10.35 p < 0.01 0.61 1 ms FW (3 , 5.55) = 53.76 p < 0.01 0.89 10 ms F (3 , 13) = 10.80 p < 0.01 0.63

U87

ANOVA Significance Effect size (Ω2) 10 ns F (4 , 19) = 2.90 p < 0.01 0.80 550 ns F (3 , 15) = 3.60 p = 0.04 0.29 1 µs F (3 , 16) = 4.32 p = 0.02 0.33 10 µs FW (4 , 9.27) = 43.54, p < 0.01 0.63 100 µs F (3 , 21) = 47.64 p < 0.01 0.85 1 ms FW (4 , 6.62) = 13.96 p < 0.01 0.71 10 ms F (3 , 19) = 10.66 p < 0.01 0.56

Undifferentiated HT22

ANOVA Significance Effect size (Ω2) 10 ns FW (3 , 3.06) = 21.88 p = 0.01 0.81 550 ns F (2 , 9) = 12.96 p < 0.01 0.67 1 µs F (3 , 9) = 11.14 p < 0.01 0.70 10 µs F (4 , 10) = 5.24 p = 0.02 0.53 100 µs F (2 , 12) = 3.79 p = 0.05 0.27 1 ms F (3 , 10) = 9.06 p < 0.01 0.63 10 ms F (4 , 13) = 4.15 p = 0.02 0.41

Differentiated HT22

ANOVA Significance Effect size (Ω2) 10 ns FW (3 , 11.76) = 13.41 p < 0.01 0.60 550 ns F (2 , 10) = 5.20 p = 0.03 0.39 1 µs F (2 , 16) = 4.56 p = 0.03 0.29 10 µs F (6 , 21) = 3.97 p < 0.01 0.39 100 µs F (4 , 16) = 6.13 p < 0.01 0.49 1 ms F (3 , 19) = 8.80 p < 0.01 0.50 10 ms F (4 , 20) = 6.02 p < 0.01 0.45

Page 108: Investigating the role of voltage-gated ion channels in

107

Table S2 - Additional statistical parameters from the strength-duration curve.

All electric fields that were tested for threshold determination are provided by cell line,

including the sample size for each.

Cell type CHO U-87 HT22 undiff HT22 diff

Control 7 11 7 8

10 ns 22 kV/m 5 16.5 kV/m 4 22 kV/m 6 44 kV/m 8

34 kV/m 3 22.0 kV/m 4 34 kV/m 4 52 kV/m 6

44 kV/m 6 34.0 kV/m 4 44 kV/m 2 76 kV/m 5

44.0 kV/m 4

550 ns 1.6 kV/cm 4 1.2 kV/cm 3 1.6 kV/cm 4 2.0 kV/cm 3

2.0 kV/cm 4 1.6 kV/cm 2 2.0 kV/cm 5 4.5 kV/cm 2

2.4 kV/cm 4 2.0 kV/cm 3

1 µs 1.0 kV/cm 4 1.0 kV/cm 3 1.0 kV/cm 4 4.5 kV/cm 2

1.2 kV/cm 4 1.2 kV/cm 3 1.2 kV/cm 3 10. kV/cm 4

1.4 kV/cm 4 1.4 kV/cm 3 1.4 kV/cm 3

10 µs 0.30 kV/cm 3 0.35 kV/m 5 0.35 kV/m 3 0.60 kV/m 3

0.45 kV/cm 4 0.45 kV/m 6 0.45 kV/m 3 0.80 kV/m 3

0.60 kV/cm 4 0.60 kV/m 4 0.60 kV/m 3 1.20 kV/m 3

0.75 kV/m 5 0.75 kV/m 3 1.60 kV/m 4

1.80 kV/m 4

2.00 kV/m 3

100 µs 0.10 kV/cm 4 0.25 kV/m 3 0.50 kV/m 3 0.35 kV/cm 3

0.20 kV/cm 4 0.35 kV/m 6 0.75 kV/m 5 0.45 kV/cm 3

0.30 kV/cm 4 0.45 kV/m 5 0.75 kV/cm 3

0.40 kV/cm 4 1.00 kV/cm 6

1 ms 0.10 kV/cm 4 0.10 kV/cm 3 0.25 kV/cm 4 0.40 kV/cm 4

0.15 kV/cm 4 0.20 kV/cm 4 0.30 kV/cm 4 0.50 kV/cm 5

0.20 kV/cm 4 0.30 kV/cm 4 0.35 kV/cm 3 0.60 kV/cm 6

0.60 kV/cm 4

10 ms 0.02 kV/cm 3 0.04 kV/cm 4 0.24 kV/cm 3 0.20 kV/cm 4

0.06 kV/cm 3 0.08 kV/cm 4 0.28 kV/cm 3 0.24 kV/cm 3

0.10 kV/cm 3 0.12 kV/cm 4 0.32 kV/cm 3 0.28 kV/cm 6

0.40 kV/cm 2 0.36 kV/cm 4

Page 109: Investigating the role of voltage-gated ion channels in

108

Chapter 5 –

Discussion and Conclusion

Page 110: Investigating the role of voltage-gated ion channels in

109

5.1 Summary

This thesis investigated the effects of pulsed electric fields on multiple cell lines. Three

chapters evaluated important questions regarding cell membrane – electric field interactions.

The first experimental chapter compared the traditional view of this interaction with

experimental results using a statistical approach. The next chapter studied the role of

transmembrane proteins in PEF effects. These proteins, or ion channels, are often overlooked

in PEF research and are critically important to a cells response to environmental stimuli.

Finally, the last chapter investigated the electric field thresholds required to depolarize the

transmembrane potential and permeabilize the plasma membrane in multiple cell lines.

Several pulse durations, ranging from 10 ns to 10 ms were used. The results from each chapter

will be discussed as they pertain to the literature.

5.2 A comparative analysis of the theoretical and experimental

interactions of PEF with cells in vitro

The effects of pulsed electric fields on biological systems has been studied for more

than a half of a century. As with any scientific inquiry, a model was required to help understand

how these fields interact with cells. But what happens when the model fails to provide a true

representation of the effects we are interested in measuring?

The results from this section indicate that the fundamental mathematical models that

were intended to help guide us in designing and interpreting experiments have very low

predictive power. Specifically, when considering size, shape, orientation to the electric field,

and density of cells exposed, the currently accepted models account for less than 5 % of the

changes observed. It is important to mention that this chapter looked at the induced

transmembrane potential as the outcome variable.

Transmembrane potential was selected because the literature states that this is the

gateway to modulating cellular activity with applied electric fields. Both older and newer

Page 111: Investigating the role of voltage-gated ion channels in

110

literature have stated that an induced transmembrane voltage of ~ 1V is required to cause

breakdown of the cell membrane [17, 18, 34, 84, 230–234]

Cell have mechanisms by which they cope with induced transmembrane voltage under

physiological conditions. Transmembrane proteins, or ion channels, control the induced

voltage through opening and closing of these proteins. These channels, although studied

mostly in excitable cells, are present to a certain degree in all cells. They are responsible for

many cell processes from cell volume regulation, DNA replication and cell division [96, 235].

5.3 Nanosecond pulsed electric fields depolarize

transmembrane potential via voltage-gated K+, Ca2+ and

TRPM8 channels in U87 glioblastoma cells

The purpose of this section was to look at electric field - cell interactions with the

following question in mind: What if the changes in transmembrane potential following a single

10 nanosecond pulsed electric field were not due to electroporation of the plasma membrane?

As we saw in the previous section, there appears to be a fundamental flaw with the description

provided by the traditional theoretical model. This is most likely due to an oversimplification

of a cell, which is not merely a leaky dielectric membrane separating two conductive solutions.

A cell is a complex, living organism that is always adapting to its environment. As a result,

modelling the cell in the traditional way has very little predictive power when analyzing

experimental results.

Since we were working with applied voltage, it seemed most logical to study cellular

mechanisms which were responsible for regulating transmembrane potential. To accomplish

this task a series of pharmacological ion channel modulators were employed to monitor their

effects on membrane depolarization following nsPEF exposure. The underlying hypothesis

supposed that if the observed change in membrane potential was due to simple diffusion of

Page 112: Investigating the role of voltage-gated ion channels in

111

ions across the electroporated plasma membrane, then we should expect no significant effect

from chemical blockers or activators of ion channels.

Although few articles have explored the possibility of PEF influencing ion channel

behavior, our results appear to be consistent with the literature. Specifically, calcium influx

was abolished following PEF exposure of bovine chromaffin cells [64], as well as GH3 and

NG108 cells [63] when specific voltage-gated calcium channel blockers were used. In the latter

study, CHO cells were used as a negative control due to the lack of voltage-gated calcium

channels. When PEF was combined with or without calcium channel blockers, no increase in

intracellular calcium was observed.

In our study we have implicated additional channels, such as voltage-gated K+ channels

and TRPM8 channels, in the depolarization response following nsPEF exposure. Although

TRPM8 channels were originally described as cold-receptors, more recently they have been

shown to display voltage sensitivity [236–238]. We had to explore the possibility that TRPM8

involvement may have been an artefact because experimentation took place at ambient

temperature. To rule this out, experiments were replicated at 37°C and similar results were

observed.

Of course the presence of voltage-gated channels will differ significantly between cell

types and possibly within the same cell type. These differences could prove to be important

when designing in vivo experiments. As an example, in certain cases electrochemotherapy

could be refined to maximize effects on a given target cell type while minimizing effects on

surrounding tissues by exploiting these differences. The following chapter has been dedicated

to investigating these effects among various cell types and pulse durations.

Page 113: Investigating the role of voltage-gated ion channels in

112

5.4 Plasma membrane depolarization and permeabilization

due to electric pulses in cell lines of different excitability

More than 20 years ago the first clinical trial used electroporation as a means to

significantly enhance chemotherapeutic uptake into tumors [239]. Since then, the study of

electrochemotherapy has continued to gain traction in the scientific and medical community.

Whether studies have utilized chemotherapeutics, high concentrations of extracellular

calcium or electric fields alone, PEF have shown to be an effective alternative to traditional

chemotherapy [40, 52, 66, 102, 182, 240, 241].

Treatment of malignant tissues with PEF has shown several significant advantages

over conventional treatments. One such advantage consists of chemical delivery directly into

the tumor. Chemotherapy is generally delivered intravenously in significant concentrations

which ultimately circulate through the entire body. Since these are highly toxic compounds,

serious secondary effects can often be expected. With electrochemotherapy,

chemotherapeutics can be injected locally to into the target tissues, and subsequent

application of PEF protocol permeabilizes the target and leads to enhanced uptake of the drug

[50, 51] while minimizing secondary effects associated with systemic administration.

Another advantage comes in the selectivity of the treatment. Whereas chemotherapy

is toxic to all cells, PEFs appear to have a greater effect on malignant cells than normal cells

[220]. Although not very well understood, this difference in sensitivity may be linked to

membrane repair. When comparing dye uptake following the standard ECT protocol and

viability 24 hours later, cancerous cell lines have been shown to be significantly more

vulnerable than normal primary cell lines [221].

Despite the advantages, one of the primary drawbacks with PEF treatment is the

potential for pain. This is due to nervous and muscular tissues in areas adjacent to the

treatment site which have become electrically excited from the applied electric field [212, 213].

The purpose of this section was to investigate PEF-induced permeability and excitation of

Page 114: Investigating the role of voltage-gated ion channels in

113

various cell lines. An important advance in PEF treatment would involve finding the

appropriate electric field duration and intensity that would minimize the excitation of

surrounding nervous tissue yet remain effective at permeabilizing the target cells.

Four cells lines were used for these experiments, three of which were non-excitable

under physiologic conditions, the other being terminally differentiated neurons. For each cell

line, PMPI was used to measure changes in membrane potential, and YO-PRO was used as a

measure of membrane permeability.

Consistent with the literature, we found that the electric field intensity required to

depolarize the transmembrane potential decreased as the pulse durations increased [25, 242,

243]. We also found that the electric field required to depolarize cells was significantly greater

in the differentiated neurons than the non-excitable cells. When we combine these results with

the fact that no difference in thresholds were observed for membrane permeability, we find

that by carefully selecting a treatment protocol with the appropriate intensity, it is possible to

achieve clinically relevant enhancement of chemical uptake while minimizing or preventing

the pain associated with excitation of the surrounding nervous tissues.

5.5 Conclusion

This thesis focused heavily on understanding the role of transmembrane ion channels

in pulsed electric field effects. These have been shown to be directly influenced in several

studies [63–65, 93]. In the first section we compared theoretical parameters influencing

membrane-electric field interactions with experimental results. We found that less than 5 %

of the experimental results could be attributed to factors included in current electro-physical

model, such as size, shape, orientation with respect to the electric field and the density of cells.

These results suggested, at least for long-lasting changes that were measured here, were

mediated to a large extent by other parameters.

Page 115: Investigating the role of voltage-gated ion channels in

114

One of those parameters are voltage-gated ion channels, which was the focus of the

second chapter. These are channels that are integral for many physiological processes and are

present, to varying degrees, in every type of cell. Although it is important that future

experiments look to understand the precise mechanism, it is no surprise that applying an

electric field would somehow influence the activity of electric field sensors in a cell. In addition

to directly acting on these channels, other research has suggested they may be damaged by

PEF exposure [224, 244, 245].

It was interesting that no ion channel blocker was able to completely abolish the

depolarizing response. It is important to consider the possibility that pores were formed in the

plasma membrane which allowed the diffusion of ions, and it was this diffusion that resulted

in the transmembrane potential being sufficiently depolarized to allow voltage-gated channels

to open. Whether or not the PEF acted directly or indirectly on the ion channels is something

that needs to be explored in the future; however, it is clear that they play a significant role in

the change in transmembrane potential following PEF exposure.

In the final experimental chapter we extended these results to include multiple cell

lines. A series of experiments looking at pulse durations from 10 ns to 10 ms were conducted

with membrane potential and membrane permeability as outcome variables. Multiple electric

field intensities were used for each duration in order to find the minimum fields strength

required to either depolarize, or permeabilize the plasma membrane. One of the most common

complaint reported from patients treated with PEF treatments is the pain associated with

repeated electrical stimulation. It would be beneficial to find a protocol which could limit or

prevent the pain associated with treatment.

Since the experienced pain is due to electrically exciting nervous tissue surrounding

the treatment site, we aimed to explore the possibility of preventing excitation without

interfering with the permeability required for enhanced drug uptake in electrochemotherapy.

Our results demonstrated that the threshold required for membrane depolarization was

Page 116: Investigating the role of voltage-gated ion channels in

115

different between cell lines. Specifically, the differentiated neurons required a significantly

greater field strength than the non-excitable cell.

In the second series of experiments investigating the threshold for

electropermeabilization in the same cells, we found that there were no significant difference

between field strengths required. This effect was consistent at every pulse duration tested.

These results are promising as they suggest it may be possible to find an electric field intensity

that would be able to enhance drug uptake, all the while avoiding excitation of the surrounding

nervous tissue associated with the sensation of pain. Most likely this effect would differ based

on several factors such as cell type, treatment site, or density of nervous tissue in the area.

Future research should focus on testing how well in vitro excitability correlates with in

vivo perception of pain. Potentially, this could reduce or eliminate the requirement for

sedatives and muscle relaxants in PEF treatments.

5.6 Perspectives

In the first experimental chapter we uncovered two significantly different populations

of U87 cells. This came in spite of the care taken to keep as many variables constant as possible,

including temperature, imaging medium, electric field intensity etc. Although there was not

enough time to experimentally pursue these differences further, these results require more

investigation. There were many interesting coincidences which may implicate cell cycle as a

target for future experimentation.

This is not something that has received a lot of attention in this field of study; however,

one study did report significant differences in PEF effects depending on the phase of cell cycle

[89]. Of course, during my last month as a Ph.D. student, an incredible tool arrived in our lab.

A Digital holographic microscope capable of monitoring cells long-term in 3D without the

need for fluorescent molecules. I truly hope future researchers will use this tool to investigate

the role of cell cycle.

Page 117: Investigating the role of voltage-gated ion channels in

116

In the second experimental chapter, we showed that by delivering a single 10 nsPEF

every 2.5 minutes, the membrane potential could not be recovered. This could be important

since cells oscillate between periods of depolarization and hyperpolarization that are

associated with DNA replication and division. If cells need to hyperpolarize to replicate their

DNA, maintaining a depolarized transmembrane potential over time could prove to be a novel

treatment for cancer that would require no chemotherapeutics and would not damage

surrounding tissues. Whether or not this is the case, I look forward to reading future results

investigating this question.

Page 118: Investigating the role of voltage-gated ion channels in

117

References

1. Barba FJ, Parniakov O, Pereira SA, et al (2015) Current applications and new

opportunities for the use of pulsed electric fields in food science and industry. Food

Res Int 77:773–798. doi: 10.1016/j.foodres.2015.09.015

2. Xue D, Farid MM (2015) Pulsed electric field extraction of valuable compounds from

white button mushroom (Agaricus bisporus). Innov Food Sci Emerg Technol 29:178–

186. doi: 10.1016/j.ifset.2015.03.012

3. Yang N, Huang K, Lyu C, Wang J (2016) Pulsed electric field technology in the

manufacturing processes of wine, beer, and rice wine: A review. Food Control 61:28–

38. doi: 10.1016/j.foodcont.2015.09.022

4. Lin L, Yahong M, Jinghui G, et al (2013) Effects of AC electric fields on the

cryopreservation of SD rat liver tissues. Annu Rep - Conf Electr Insul Dielectr

Phenomena, CEIDP 555–558. doi: 10.1109/CEIDP.2013.6748328

5. Dovgan B, Barlič A, Knežević M, Miklavčič D (2016) Cryopreservation of Human

Adipose-Derived Stem Cells in Combination with Trehalose and Reversible

Electroporation. J Membr Biol. doi: 10.1007/978-981-287-817-5

6. Dovgan B, Dermol J, Barlič A, et al (2016) Cryopreservation of Human Umbilical Stem

Cells in Combination with Trehalose and Reversible Electroporation. In: Jarm T,

Kramar P (eds) 1st World Congr. Electroporation Pulsed Electr. Fields Biol. Med.

Food {&} Environ. Technol. Portoro{ž}, Slov. Sept. 6 --10, 2015. Springer Singapore,

Singapore, pp 307–310

7. Cemazar M, Sersa G (2007) Electrotransfer of therapeutic molecules into tissues. Curr

Opin Mol Ther 9:554–562.

8. Pucihar G, Kotnik T, Miklavcic D, Teissie J (2008) Kinetics of transmembrane

transport of small molecules into electropermeabilized cells. Biophys J 95:2837–

2848. doi: 10.1529/biophysj.108.135541

Page 119: Investigating the role of voltage-gated ion channels in

118

9. Calvet CY, Thalmensi J, Liard C, et al (2014) Optimization of a gene electrotransfer

procedure for efficient intradermal immunization with an hTERT-based DNA vaccine

in mice. Mol Ther Methods Clin Dev 1:14045. doi: 10.1038/mtm.2014.45

10. Chopinet L, Batista-Napotnik T, Montigny A, et al (2013) Nanosecond electric pulse

effects on gene expression. J Membr Biol 246:851–859. doi: 10.1007/s00232-013-

9579-y

11. Pakhomova ON, Gregory BW, Semenov I, Pakhomov AG (2013) Two Modes of Cell

Death Caused by Exposure to Nanosecond Pulsed Electric Field. PLoS One. doi:

10.1371/journal.pone.0070278

12. Beebe SJ, Sain NM, Ren W (2013) Induction of Cell Death Mechanisms and Apoptosis

by Nanosecond Pulsed Electric Fields (nsPEFs). Cells 2:136–62. doi:

10.3390/cells2010136

13. Rubinsky B, Onik G, Mikus P (2007) Irreversible electroporation: A new ablation

modality - Clinical implications. Technol Cancer Res Treat 6:37–48. doi:

10.1177/153303460700600106

14. Meer G Van, Voelker DR, Feigenson GW (2008) Membrane lipids : where they are

and how they behave. Nat Rev Mol cell Biol 9:112–124. doi: 10.1038/nrm2330

15. Alberts B, Johnson A, Lewis J, et al (2002) Molecular biology of the cell, 4th ed.

Garland Science

16. Tortora G, Derrickson B (2008) Principles of anatomy and physiology, 12th ed. Wiley

17. Teissie J, Golzio M, Rols MP (2005) Mechanisms of cell membrane

electropermeabilization: A minireview of our present (lack of ?) knowledge. Biochim

Biophys Acta - Gen Subj 1724:270–280. doi: 10.1016/j.bbagen.2005.05.006

18. Rems L, Miklavcic D (2016) Tutorial: Electroporation of cells in complex materials

and tissue. J Appl Phys. doi: 10.1063/1.4949264

Page 120: Investigating the role of voltage-gated ion channels in

119

19. Gurtovenko AA, Vattulainen I (2005) Pore Formation Coupled to Ion Transport

through Lipid Membranes as Induced by Transmembrane Ionic Charge Imbalance :

Atomistic Molecular Dynamics Study. J Am Chem Soc 127:17570–17571.

20. Tarek M (2005) Membrane Electroporation: A Molecular Dynamics Simulation.

Biophys J 88:4045–4053. doi: 10.1529/biophysj.104.050617

21. Tieleman DP (2004) The molecular basis of electroporation. BMC Biochem 5:10. doi:

10.1186/1471-2091-5-10

22. Jianfei WJW, Yan MYM, Chenguo YCY, Chengxiang LCL (2008) A Review on

Molecular Dynamics Simulation for Biological Effects of Pulsed Electric Field. 2008

Int Conf High Volt Eng Appl 763–766. doi: 10.1109/ICHVE.2008.4774046

23. Vernier PT, Ziegler MJ (2007) Nanosecond Field Alignment of Head Group and Water

Dipoles in Electroporating Phospholipid Bilayers. J Phys Chem 111:12993–12996.

24. Sözer EB, Levine ZA, Vernier PT (2017) Quantitative Limits on Small Molecule

Transport via the Electropermeome — Measuring and Modeling Single Nanosecond

Perturbations. Sci Rep 7:57. doi: 10.1038/s41598-017-00092-0

25. Rogers WR, Merritt JH, Comeaux JA, et al (2004) Strength-duration curve an

electrically excitable tissue extended down to near 1 nanosecond. IEEE Trans Plasma

Sci 32:1587–1599. doi: 10.1109/TPS.2004.831758

26. Romeo S, Wu YH, Levine Z a., et al (2013) Water influx and cell swelling after

nanosecond electropermeabilization. Biochim Biophys Acta - Biomembr 1828:1715–

1722. doi: 10.1016/j.bbamem.2013.03.007

27. Carr L, Bardet SM, Burke RC, et al (2017) Calcium-independent disruption of

microtubule dynamics by nanosecond pulsed electric fields in U87 human

glioblastoma cells. Sci Rep. doi: 10.1038/srep41267

28. Kotnik T, Mir LM, Flisar K, et al (2001) Cell membrane electropermeabilization by

Page 121: Investigating the role of voltage-gated ion channels in

120

symmetrical bipolar rectangular pulses: Part I. Increased efficiency of

permeabilization. Bioelectrochemistry 54:83–90.

29. Pakhomov AG, Semenov I, Xiao S, et al (2014) Cancellation of cellular responses to

nanoelectroporation by reversing the stimulus polarity. Cell Mol Life Sci 71:4431–

4441. doi: 10.1007/s00018-014-1626-z

30. Schoenbach KH, Pakhomov AG, Semenov I, et al (2014) Ion transport into cells

exposed to monopolar and bipolar nanosecond pulses. Bioelectrochemistry. doi:

10.1016/j.bioelechem.2014.08.015

31. Ibey BL, Ullery J, Pakhomova ON, et al (2014) Bipolar nanosecond electric pulses are

less efficient at electropermeabilization and killing cells than monopolar pulses.

Biochem Biophys Res Commun 443:568–573. doi:

10.1016/j.bbrc.2013.12.004.Bipolar

32. Gowrishankar TR, Weaver JC (2006) Electrical behavior and pore accumulation in a

multicellular model for conventional and supra-electroporation. Biochem Biophys Res

Commun 349:643–653. doi: 10.1016/j.bbrc.2006.08.097

33. Tekle E, Oubrahim H, Dzekunov SM, et al (2005) Selective Field Effects on

Intracellular Vacuoles and Vesicle Membranes with Nanosecond Electric Pulses.

Biophys J 89:274–284. doi: 10.1529/biophysj.104.054494

34. Schwan HP (1988) Biological effects of non-ionizing radiations: Cellular properties

and interactions. Ann Biomed Eng 16:245–263. doi: 10.1007/BF02368002

35. Dev SB, Rabussay DP, Widera G, Hofmann GA (2000) Medical Applications of

Electroporation. IEEE Trans plasma Sci 28:206–223.

36. Vandermeulen G, Richiardi H, Escriou V, et al (2009) Skin-specific promoters for

genetic immunisation by DNA electroporation. Vaccine 27:4272–4277. doi:

10.1016/j.vaccine.2009.05.022

Page 122: Investigating the role of voltage-gated ion channels in

121

37. Heller R, Cruz Y, Heller LC, et al (2010) Electrically Mediated Delivery of Plasmid

DNA to the Skin, Using a Multielectrode Array. Hum Gene Ther 21:357–362. doi:

10.1089/hum.2009.065

38. Teissié J (2013) Electrically Mediated Gene Delivery: Basic and Translational

Concepts. Nov Gene Ther Approaches. doi: 10.5772/54780

39. Gothelf A, Gehl J (2010) Gene electrotransfer to skin; review of existing literature and

clinical perspectives. Curr Gene Ther 10:287–299. doi: 10.2174/156652310791823443

40. Gothelf A, Mir LM, Gehl J (2003) Electrochemotherapy : results of cancer treatment

using enhanced delivery of bleomycin by electroporation. Cancer Treat Rev 29:371–

387. doi: 10.1016/S0305-7372(03)00073-2

41. Breton M, Mir LM (2012) Microsecond and nanosecond electric pulses in cancer

treatments. Bioelectromagnetics 33:106–123. doi: 10.1002/bem.20692

42. Davalos R V., Mir LM, Rubinsky B (2005) Tissue Ablation with Irreversible

Electroporation. Ann Biomed Eng 33:223–231. doi: 10.1007/s10439-005-8981-8

43. Ting F, Tran M, Böhm M, et al (2016) Focal irreversible electroporation for prostate

cancer: functional outcomes and short-term oncological control. Prostate Cancer

Prostatic Dis 19:46–52. doi: 10.1038/pcan.2015.47

44. Rossmeisl JH, Garcia PA, Pancotto TE, et al (2015) Safety and feasibility of the

NanoKnife system for irreversible electroporation ablative treatment of canine

spontaneous intracranial gliomas. J Neurosurg 123:1008–1025. doi:

10.3171/2014.12.JNS141768

45. Jiang C, Davalos R V., Bischof JC (2015) A Review of Basic to Clinical Studies of

Irreversible Electroporation Therapy. IEEE Trans Biomed Eng 62:4–20. doi:

10.1109/TBME.2014.2367543

46. Scheffer HJ, Nielsen K, de Jong MC, et al (2014) Irreversible Electroporation for

Page 123: Investigating the role of voltage-gated ion channels in

122

Nonthermal Tumor Ablation in the Clinical Setting: A Systematic Review of Safety and

Efficacy. J Vasc Interv Radiol 1–15. doi: 10.1016/j.jvir.2014.01.028

47. Fruhling P, Nilsson A, Duraj F, et al (2017) Single-center nonrandomized clinical trial

to assess the safety and efficacy of irreversible electroporation ( IRE ) ablation of liver

tumors in humans : Short to mid-term results. Eur J Surg Oncol 1–7. doi:

10.1016/j.ejso.2016.12.004

48. Heller R, Heller LC (2015) Gene Electrotransfer Clinical Trials. In: Adv. Genet.

Elsevier, pp 235–262

49. Orlowski S, Belehradek JJ, Paoletti C, Mir L (1988) Transient electropermeabilization

of cells in culture. Increase of the cytotoxicity of anticancer drugs. Biochem Pharmacol

37:4727–4733.

50. Horiuchi A, Nikaido T, Mitsushita J, et al (2000) Enhancement of antitumor effect of

bleomycin by low-voltage in vivo electroporation- a study of human uterine

leiomyosarcomas in nude mice. Int J cancer J 88:640–644.

51. Gehl J, Skovsgaard T, Mir LM (1998) Enhancement of cytotoxicity by

electropermeabilization: an improved method for screening drugs. Anticancer drugs

9:319–325.

52. Falk H, Lambaa S, Johannesen HH, et al (2017) Electrochemotherapy and calcium

electroporation inducing a systemic immune response with local and distant

remission of tumors in a patient with malignant melanoma – a case report. Acta Oncol

(Madr). doi: 10.1080/0284186X.2017.1290274

53. Beebe SJ, Chen YJ, Sain NM, et al (2012) Transient Features in Nanosecond Pulsed

Electric Fields Differentially Modulate Mitochondria and Viability. PLoS One. doi:

10.1371/journal.pone.0051349

54. Ren W, Beebe SJ (2011) An apoptosis targeted stimulus with nanosecond pulsed

Page 124: Investigating the role of voltage-gated ion channels in

123

electric fields (nsPEFs) in E4 squamous cell carcinoma. Apoptosis 16:382–393. doi:

10.1007/s10495-010-0572-y

55. Beebe SJ, Fox PM, Rec LJ, et al (2002) Nanosecond Pulsed Electric Field ( nsPEF )

Effects on Cells and Tissues : Apoptosis Induction and Tumor Growth Inhibition.

IEEE Trans plasma Sci 30:286–292.

56. Vernier PT, Li A, Marcu L, et al (2003) Ultrashort Pulsed Electric Fields Induce

Membrane Phospholipid Translocation and Caspase Activation : Differential

Sensitivities of Jurkat T Lymphoblasts and Rat Glioma C6 Cells. IEEE Trans Dielectr

Electr Insul 10:795–809.

57. Chen N, Schoenbach KH, Kolb JF, et al (2004) Leukemic cell intracellular responses

to nanosecond electric fields. Biochem Biophys Res Commun 317:421–427. doi:

10.1016/j.bbrc.2004.03.063

58. Yin S, Chen X, Hu C, et al (2014) Nanosecond pulsed electric field (nsPEF) treatment

for hepatocellular carcinoma : A novel locoregional ablation decreasing lung

metastasis. Cancer Lett 346:285–291. doi: 10.1016/j.canlet.2014.01.009

59. Stacey M, Fox P, Buescher S, Kolb J (2011) Nanosecond pulsed electric field induced

cytoskeleton, nuclear membrane and telomere damage adversely impact cell survival.

Bioelectrochemistry 82:131–134. doi: 10.1016/j.bioelechem.2011.06.002

60. Silve A, Leray I, Mir LM (2012) Demonstration of cell membrane permeabilization to

medium-sized molecules caused by a single 10 ns electric pulse. Bioelectrochemistry

87:260–264. doi: 10.1016/j.bioelechem.2011.10.002

61. Breton M, Delemotte L, Silve A, et al (2012) Transport of siRNA through Lipid

Membranes Driven by Nanosecond Electric Pulses: An Experimental and

Computational Study. J Am Chem Soc 134:13938–13941.

62. Gowrishankar, Thiruvallur, R Esser AT, Vasilkoski Z, Smith KC, Weaver JC (2006)

Page 125: Investigating the role of voltage-gated ion channels in

124

Microdosimetry for conventional and supra-electroporation in cells with organelles.

Biochem Biophys Res Commun 341:1266–1276. doi: 10.1016/j.bbrc.2006.01.094

63. Semenov I, Xiao S, Kang D, et al (2015) Cell stimulation and calcium mobilization by

picosecond electric pulses. Bioelectrochemistry 105:65–71. doi:

10.1016/j.bioelechem.2015.05.013

64. Craviso GL, Choe S, Chatterjee P, et al (2010) Nanosecond electric pulses: A novel

stimulus for triggering Ca2+ influx into chromaffin cells via voltage-gated Ca2+

channels. Cell Mol Neurobiol 30:1259–1265. doi: 10.1007/s10571-010-9573-1

65. Pakhomov AG, Semenov I, Casciola M, Xiao S (2017) Neuronal excitation and

permeabilization by 200-ns pulsed electric field: An optical membrane potential study

with FluoVolt dye. BBA - Biomembr. doi: 10.1016/j.bbamem.2017.04.016

66. Nuccitelli R, Wood R, Kreis M, et al (2014) First-in-human trial of

nanoelectroablation therapy for basal cell carcinoma : proof of method. Exp Dermatol

23:130–142. doi: 10.1111/exd.12303

67. Miklavčič D (2017) Handbook of electroporation. Springer International Publishing

68. Mazères S, Sel D, Golzio M, et al (2009) Non invasive contact electrodes for in vivo

localized cutaneous electropulsation and associated drug and nucleic acid delivery. J

Control Release 134:125–131. doi: 10.1016/j.jconrel.2008.11.003

69. Vernier PT, Sun Y, Gundersen MA (2006) Nanoelectropulse-driven membrane

perturbation and small molecule permeabilization. BMC Cell Biol 7:1–16. doi:

10.1186/1471-2121-7-37

70. Jin L, Platisa J, Wooltorton JRA, et al (2012) Single action potentials and

subthreshold electrical events imaged in neurons with a novel fluorescent protein

voltage probe. Neuron 75:779–785. doi: 10.1016/j.neuron.2012.06.040.Single

71. Boyce FM, Bucher NLR (1996) Baculovirus-mediated transfer into mammalian cells.

Page 126: Investigating the role of voltage-gated ion channels in

125

Proc Natl Acad Sci 93:2348–2352.

72. Ghosh S, Parvez K, Banerjee K, et al (2002) Baculovirus as Mammalian Cell

Expression Vector for Gene Therapy : An Emerging Strategy. Mol Ther 6:5–11. doi:

10.1006/mthe.2000.0643

73. Kost TA, Condreat JP, Jarvis DL (2005) Baculovirus as versatile vectors for protein

expression in insect and mammalian cells. Nat Biotechnol 23:567–575. doi:

10.1038/nbt1095.Baculovirus

74. Scaduto RC, Grotyohann LW (1999) Measurement of mitochondrial membrane

potential using fluorescent rhodamine derivatives. Biophys J 76:469–477. doi:

10.1016/S0006-3495(99)77214-0

75. Solaini G, Sgarbi G, Lenaz G, Baracca A (2007) Evaluating Mitochondrial Membrane

Potential in Cells. Biosci Rep 27:11–21. doi: 10.1007/s10540-007-9033-4

76. Plfigek J, Sigler K (1996) Slow fluorescent indicators of membrane potential : a survey

of different approaches to probe response analysis. J Photochem Photobiol B Biol

33:101–124.

77. Nicholls DG (2006) Simultaneous monitoring of ionophore- and inhibitor-mediated

plasma and mitochondrial membrane potential changes in cultured neurons. J Biol

Chem 281:14864–74. doi: 10.1074/jbc.M510916200

78. Spitzner M, Ousingsawat J, Scheidt K, et al (2007) Voltage-gated K+ channels support

proliferation of colonic carcinoma cells. FASEB J 21:35–44. doi: 10.1096/fj.06-

6200com

79. Whiteaker KL, Gopalakrishnan SM, Groebe D, et al (2001) Validation of FLIPR

membrane potential dye for high throughput screening of potassium channel

modulators. J Biomol Screen 6:305–312. doi: 0803973233

80. Dorn A, Hermann F, Ebneth A, et al (2005) Evaluation of a High-Throughput

Page 127: Investigating the role of voltage-gated ion channels in

126

Fluorescence Assay Method for hERG Potassium Channel Inhibition. J Biomol Screen

10:339–347. doi: 10.1177/1087057104272045

81. Field A (2009) Discovering statistics using SPSS, 3rd ed. SAGE publications Ltd

82. Schwan HP (1957) Electrical properties of tissue and cell suspensions. Adv Biol Med

Phys 5:147–209.

83. Gabriel B, Teissie J (1999) Time Courses of Mammalian Cell Electropermeabilization

Observed by Millisecond Imaging of Membrane Property Changes during the Pulse.

Biophys J 76:2158–2165.

84. Frey W, White J a, Price RO, et al (2006) Plasma membrane voltage changes during

nanosecond pulsed electric field exposure. Biophys J 90:3608–3615. doi:

10.1529/biophysj.105.072777

85. Valic B, Golzio M, Pavlin M, et al (2003) Electric field induced trans-membrane

potential on spheroidal cell : theory and experiment. Eur Biophys J 32:519–528. doi:

10.1007/s00249-003-0296-9

86. Kotnik T, Pucihar G, Miklavcic D (2010) Induced Transmembrane Voltage and Its

Correlation with Electroporation-Mediated Molecular Transport. J Membr Biol

236:3–13. doi: 10.1007/s00232-010-9279-9

87. Carr L, Bardet SM, Arnaud-cormos D, et al (2018) Visualisation of an nsPEF induced

calcium wave using the genetically encoded calcium indicator GCaMP in U87 human

glioblastoma cells. Bioelectrochemistry 119:68–75. doi:

10.1016/j.bioelechem.2017.09.003

88. Puc M, Kotnik T, Mir LM, Miklavc D (2003) Quantitative model of small molecules

uptake after in vitro cell electropermeabilization. Bioelectrochemistry 60:1–10. doi:

10.1016/S1567-5394(03)00021-5

89. Hojo S, Shimizu K, Yositake H, et al (2003) The Relationship Between

Page 128: Investigating the role of voltage-gated ion channels in

127

Electropermeabilization and Cell Cycle and Cell Size of Saccharomyces Cerevisiae.

IEEE Trans Nanobioscience 2:35–39.

90. Sukhorukov VL, Djuzenova CS, Frank H, et al (1995) Electropermeabilization and

Fluorescent Tracer Exchange : The Role of Whole-Cell Capacitance. Cytometry

21:230–240.

91. Henslee BE, Morss A, Hu X, et al (2011) Electroporation Dependence on Cell Size :

Optical Tweezers Study. Anal Chem 83:3998–4003.

92. Denzi A, Camera F, Merla C, et al (2016) A Microdosimeric Study of Electropulsation

on Multiple Realistically Shaped Cells : Effect of Neighbours. J Membr Biol. doi:

10.1007/s00232-016-9912-3

93. Teissie J, Tsong TY (1980) Evidence of voltage-induced channel opening in Na/K

ATPase of human erythrocyte membrane. J Membr Biol 55:133–140. doi:

10.1007/BF01871155

94. Chen W, Lee RC (1994) Altered ion channel conductance and ionic selectivity induced

by large imposed membrane potential pulse. Biophys J 67:603–612. doi:

10.1016/S0006-3495(94)80520-X

95. Levin M (2007) Large-scale biophysics: ion flows and regeneration. Trends Cell Biol

17:261–270. doi: 10.1016/j.tcb.2007.04.007

96. Blackiston DJ, McLaughlin KA, Levin M (2009) Bioelectric controls of cell

proliferation: Ion channels, membrane voltage and the cell cycle. Cell Cycle 8:3519–

3528. doi: 10.1115/1.3071969.Automating

97. Rao VR, Perez-Neut M, Kaja S, Gentile S (2015) Voltage-gated ion channels in cancer

cell proliferation. Cancers (Basel) 7:849–875. doi: 10.3390/cancers7020813

98. Tay D, Bhathal P, Fox R (1991) Quantitation of Go and G1 Phase Cells in Primary

Carcinomas. J Clin Invest 87:519–527.

Page 129: Investigating the role of voltage-gated ion channels in

128

99. Motaln H, Koren A, Gruden K, et al (2015) Heterogeneous glioblastoma cell cross-talk

promotes phenotype alterations and enhanced drug resistance. Oncotarget 6:40998–

41017.

100. Burke RC, Bardet SM, Carr L, et al (2017) Nanosecond pulsed electric fields depolarize

transmembrane potential via voltage-gated K+, Ca2+ and TRPM8 channels in U87

glioblastoma cells. BBA - Biomembr 1859:2040–2050. doi:

10.1016/j.bbamem.2017.07.004

101. Esser AT, Smith KC, Gowrishankar TR, Weaver JC (2009) Towards Solid Tumor

Treatment by Nanosecond Pulsed Electric Fields. Technol cancer Res Treat 8:289–

306.

102. Nuccitelli R, Pliquett U, Chen X, et al (2006) Nanosecond pulsed electric fields cause

melanomas to self-destruct. Biochem Biophys Res Commun 343:351–360. doi:

10.1016/j.bbrc.2006.02.181

103. Guo S, Jackson DL, Burcus NI, et al (2014) Gene electrotransfer enhanced by

nanosecond pulsed electric fields. Mol Ther Methods Clin Dev 1:14043. doi:

10.1038/mtm.2014.43

104. SJ B, J W, PF B, et al (2003) Diverse effects of nanosecond pulsed electric fields on

cells and tissues. DNA Cell Biol 22:785–796.

105. Pakhomov AG, Kolb JF, White J a., et al (2007) Long-lasting plasma membrane

permeabilization in mammalian cells by nanosecond Pulsed Electric Field (nsPEF).

Bioelectromagnetics 28:655–663. doi: 10.1002/bem.20354

106. Cantu JC, Tarango M, Beier HT, Ibey BL (2016) The Biological Response Of Cells To

Nanosecond Pulsed Electric Fields Is Dependent On Plasma Membrane Cholesterol.

Biochim Biophys Acta - Biomembr 1858:2636–2646. doi:

10.1016/j.bbamem.2016.07.006

Page 130: Investigating the role of voltage-gated ion channels in

129

107. Semenov I, Xiao S, Pakhomov AG (2013) Primary pathways of intracellular Ca(2+)

mobilization by nanosecond pulsed electric field. Biochim Biophys Acta 1828:981–9.

doi: 10.1016/j.bbamem.2012.11.032

108. Son RS, Smith KC, Gowrishankar TR, et al (2014) Basic Features of a Cell

Electroporation Model : Illustrative Behavior for Two Very Different Pulses. J Membr

Biol 247:1209–1228. doi: 10.1007/s00232-014-9699-z

109. Bardet SM, Carr L, Soueid M, et al (2016) Multiphoton imaging reveals that

nanosecond pulsed electric fields collapse tumor and normal vascular perfusion in

human glioblastoma xenografts. Sci Rep 6:1–11. doi: 10.1038/srep34443

110. Tokman M, Lee JH, Levine ZA, et al (2013) Electric Field-Driven Water Dipoles:

Nanoscale Architecture of Electroporation. PLoS One. doi:

10.1371/journal.pone.0061111

111. Bowman AM, Nesin OM, Pakhomova ON, Pakhomov AG (2010) Analysis of plasma

membrane integrity by fluorescent detection of Tl + uptake. J Membr Biol 236:15–26.

doi: 10.1007/s00232-010-9269-y

112. Pakhomov AG, Bowman AM, Ibey BL, et al (2009) Lipid Nanopores Can Form a

Stable, Ion Channel-Like Conduction Pathway in Cell Membrane. Biochem Biophys

Res Commun 385:181–186. doi: 10.1016/j.bbrc.2009.05.035.Lipid

113. Li M, Xiong ZG (2011) Ion channels as targets for cancer therapy. Int J Physiol

Pathophysiol Pharmacol 3:156–166. doi: 10.1115/1.4026364

114. Prevarskaya N, Skryma R, Shuba Y (2010) Ion channels and the hallmarks of cancer.

Trends Mol Med 16:107–121. doi: 10.1016/j.molmed.2010.01.005

115. Clark MJ, Homer N, O’Connor BD, et al (2010) U87MG decoded: The genomic

sequence of a cytogenetically aberrant human cancer cell line. PLoS Genet. doi:

10.1371/journal.pgen.1000832

Page 131: Investigating the role of voltage-gated ion channels in

130

116. Ducret T, Vacher A-M, Vacher P (2003) Voltage-dependent ionic conductances in the

human malignant astrocytoma cell line U87-MG. Mol Membr Biol 20:329–343. doi:

10.1080/0968763031000138037

117. Molenaar RJ (2011) Ion Channels in Glioblastoma. ISRN Neurol 2011:1–7. doi:

10.5402/2011/590249

118. Chin LS, Park CC, Zitnay KM, et al (1997) 4-Aminopyridine Causes Apoptosis and

Blocks an Outward Rectifier K + Channel in Malignant Astrocytoma Cell Lines. J

Neurosci Res 48:122–127.

119. Chattopadhyay N, Ye C-P, Yamaguchi T, et al (1999) Evidence for Extracellular

Calcium-Sensing Receptor Mediated Opening of an Outward K ؉ Channel in a Human

Astrocytoma Cell Line ( U87 ). Glia 26:64–72.

120. McFerrin MB, Sontheimer H (2006) A role for ion channels in glioma cell invasion.

Neuron Glia Biol 2:39–49. doi: 10.1017/S17440925X06000044

121. Arcangeli A, Crociani O, Lastraioli E, et al (2009) Targeting Ion Channels in Cancer :

A Novel Frontier in Antineoplastic Therapy. Curr Med Chem 16:66–93.

122. Delmas P, Gola M (1995) Choline blocks large conductance Kca channels in

mammalian sympathetic neurones. Neurosci Lett 189:109–112.

123. Lippiat JD, Standen NB, Davies NW (1998) Block of cloned BK(Ca) channels (rSlo)

expressed in HEK 293 cells by N-methyl D-glucamine. Pflugers Arch Eur J Physiol

436:810–812. doi: 10.1007/s004240050708

124. Wolff C, Fuks B, Chatelain P (2003) Comparative Study of Membrane Potential-

Sensitive Fluorescent Probes and their Use in Ion Channel Screening Assays. J Biomol

Screen 8:533–543. doi: 10.1177/1087057103257806

125. Goehring I, Gerencser AA, Schmidt S, et al (2012) Plasma membrane potential

oscillations in insulin secreting Ins-1 832/13 cells do not require glycolysis and are not

Page 132: Investigating the role of voltage-gated ion channels in

131

initiated by fluctuations in mitochondrial bioenergetics. J Biol Chem 287:15706–

15717. doi: 10.1074/jbc.M111.314567

126. Baxter DF, Kirk M, Garcia AF, et al (2002) A novel membrane potential-sensitive

fluorescent dye improves cell-based assays for ion channels. J Biomol Screen Off J

Soc Biomol Screen 7:79–85. doi: 10.1177/108705710200700110

127. Yan J, Aldrich RW (2010) LRRC26 auxiliary protein allows BK channel activation at

resting voltage without calcium. Nature 466:513–516. doi: 10.1038/nature09162

128. Warbington L, Hillman T, Adams C, Stern M (1996) Reduced transmitter release

conferred by mutations in the slowpoke-encoded Ca2+-activated K+ channel gene of

Drosophila. Invertebr Neurosci 2:51–60.

129. Pattillo JM, Yazejian B, Digregorio DA, et al (2001) Contribution of presynaptic

calcium-activated potassium currents to transmitter release regulation in cultured

xenopus nerve-muscle synapses. Neuroscience 102:229–240.

130. Skinner LJ, Enée V, Beurg M, et al (2003) Contribution of BK Ca2+ -Activated K+

Channels to Auditory Neurotransmission in the Guinea Pig Cochlea. J Neurophysiol

90:320–332.

131. Xu JW, Slaughter MM (2005) Large-Conductance Calcium-Activated Potassium

Channels Facilitate Transmitter Release in Salamander Rod Synapse. J Neurosci

25:7660–7668. doi: 10.1523/JNEUROSCI.1572-05.2005

132. Guéguinou M, Chantôme A, Fromont G, et al (2014) KCa and Ca2+ channels: The

complex thought. Biochim Biophys Acta 1843:2322–2333. doi:

10.1016/j.bbamcr.2014.02.019

133. Schütz GJ, Kada G, Pastushenko VP, Schindler H (2000) Properties of lipid

microdomains in a muscle cell membrane visualized by single molecule microscopy.

EMBO J 19:892–901. doi: 10.1093/emboj/19.5.892

Page 133: Investigating the role of voltage-gated ion channels in

132

134. Dart C (2010) Lipid microdomains and the regulation of ion channel function. J

Physiol 588:3169–78. doi: 10.1113/jphysiol.2010.191585

135. Gackière F, Warnier M, Katsogiannou M, et al (2013) Functional coupling between

large-conductance potassium channels and Cav3.2 voltage-dependent calcium

channels participates in prostate cancer cell growth. Biol Open 2:941–51. doi:

10.1242/bio.20135215

136. Grunnet M, Kaufmann WA (2004) Coassembly of big conductance Ca2+-activated K+

channels and L-type voltage-gated Ca2+ channels in rat brain. J Biol Chem

279:36445–36453. doi: 10.1074/jbc.M402254200

137. Wondergem R, Bartley JW (2009) Menthol increases human glioblastoma

intracellular Ca2+, BK channel activity and cell migration. J Biomed Sci 16:1–7. doi:

10.1186/1423-0127-16-90

138. Reinhart PH, Chung S, Levitan IB (1989) A family of calcium-dependent potassium

channels from rat brain. Neuron 2:1031–1041. doi: 0896-6273(89)90227-4 [pii]

139. Benton MD, Lewis AH, Bant JS, Raman IM (2013) Iberiotoxin-sensitive and -

insensitive BK currents in Purkinje neuron somata. J Neurophysiol 109:2528–41. doi:

10.1152/jn.00127.2012

140. Yu M, Liu S, Sun P, et al (2016) Peptide toxins and small-molecule blockers of BK

channels. Acta Pharmacol Sin 37:56–66. doi: 10.1038/aps.2015.139

141. Berkefeld H, Fakler B, Schulte UWE (2010) Ca 2+-Activated K+ Channels : From

Protein Complexes to Function. Physiol Rev 90:1437–1459. doi:

10.1152/physrev.00049.2009.

142. Wang B, Jaffe DB, Brenner R (2014) Current understanding of iberiotoxin-resistant

BK channels in the nervous system. Front Physiol 5:1–11. doi:

10.3389/fphys.2014.00382

Page 134: Investigating the role of voltage-gated ion channels in

133

143. Shruti S, Urban-Ciecko J, Fitzpatrick JA, et al (2012) The brain-specific beta4 subunit

downregulates BK channel cell surface expression. PLoS One. doi:

10.1371/journal.pone.0033429

144. Meera P, Wallner M, Toro L (2000) A neuronal beta subunit (KCNMB4) makes the

large conductance, voltage- and Ca2+-activated K+ channel resistant to charybdotoxin

and iberiotoxin. Proc Natl Acad Sci U S A 97:5562–7. doi: 10.1073/pnas.100118597

145. Liu X, Chang Y, Reinhart PH, et al (2002) Cloning and characterization of glioma BK,

a novel BK channel isoform highly expressed in human glioma cells. J Neurosci

22:1840–9. doi: 22/5/1840 [pii]

146. Marie C, Verkerke HP, Theodorescu D, Petri WA (2015) A whole-genome RNAi screen

uncovers a novel role for human potassium channels in cell killing by the parasite

Entamoeba histolytica. Sci Rep 5:1–18. doi: 10.1038/srep13613\rsrep13613 [pii]

147. Brelidze TI, Magleby KL (2004) Protons block BK channels by competitive inhibition

with K+ and contribute to the limits of unitary currents at high voltages. J Gen Physiol

123:305–19. doi: 10.1085/jgp.200308951

148. Peier AM, Moqrich A, Hergarden AC, et al (2002) A TRP Channel that Senses Cold

Stimuli and Menthol. Cell 108:705–715. doi: 10.1016/S0092-8674(02)00652-9

149. Chen J, Kim D, Bianchi BR, et al (2009) Pore dilation occurs in TRPA1 but not in

TRPM8 channels. Mol Pain 5:1–6. doi: 10.1186/1744-8069-5-3

150. Harteneck C (2005) Function and pharmacology of TRPM cation channels. Naunyn

Schmiedebergs Arch Pharmacol 371:307–314. doi: 10.1007/s00210-005-1034-x

151. Tominaga M, Caterina MJ (2004) Thermosensation and pain. J Neurobiol 61:3–12.

doi: 10.1002/neu.20079

152. Yoon J, Leblanc N, Zaklit J, et al (2016) Enhanced Monitoring of Nanosecond Electric

Pulse-Evoked Membrane Conductance Changes in Whole-Cell Patch Clamp

Page 135: Investigating the role of voltage-gated ion channels in

134

Experiments. J Membr Biol 249:1–12. doi: 10.1007/s00232-016-9902-5

153. Binshtok AM, Bean BP, Woolf CJ (2007) Inhibition of nociceptors by TRPV1-

mediated entry of impermeant sodium channel blockers. Nature 449:607–10. doi:

10.1038/nature06191

154. Banke TG, Chaplan SR, Wickenden a D (2010) Dynamic changes in the TRPA1

selectivity filter lead to progressive but reversible pore dilation. Am J Physiol Cell

Physiol 298:C1457-68. doi: 10.1152/ajpcell.00489.2009

155. Cao E, Liao M, Cheng Y, Julius D (2013) TRPV1 structures in distinct conformations

reveal mechanisms of activation. Nature 504:113–118. doi: 10.1038/nbt.3121.ChIP-

nexus

156. Chung M-K, Güler AD, Caterina MJ (2008) TRPV1 shows dynamic ionic selectivity

during agonist stimulation. Nat Neurosci 11:555–564. doi: 10.1038/nn.2102

157. Cui J, Yang H, Lee US (2009) Molecular Mechanisms of BK Channel Activation. Cell

Mol life Sci 66:852–875. doi: 10.1007/s00018-008-8609-x.Molecular

158. Ge L, Hoa NT, Wilson Z, et al (2014) Big Potassium (BK) ion channels in biology,

disease and possible targets for cancer immunotherapy. Int Immunopharmacol

22:427–443. doi: 10.1016/j.intimp.2014.06.040

159. Ge L, Hoa NT, Cornforth AN, et al (2012) Glioma big potassium channel expression in

human cancers and possible T cell epitopes for their immunotherapy. J Immunol

189:2625–2634. doi: 10.4049/jimmunol.1102965

160. Han X, Xi L, Wang H, et al (2008) The potassium ion channel opener NS1619 inhibits

proliferation and induces apoptosis in A2780 ovarian cancer cells. Biochem Biophys

Res Commun 375:205–209. doi: 10.1016/j.bbrc.2008.07.161

161. Cox DH, Cui J, Aldrich RW (1997) Allosteric gating of a large conductance Ca-

activated K+ channel. J Gen Physiol 110:257–281. doi: 10.1085/jgp.110.3.257

Page 136: Investigating the role of voltage-gated ion channels in

135

162. Zhao H, Sokabe M (2008) Tuning the mechanosensitivity of a BK channel by changing

the linker length. Cell Res 18:871–8. doi: 10.1038/cr.2008.88

163. Roth CC, Barnes R a, Ibey BL, et al (2015) Characterization of Pressure Transients

Generated by Nanosecond Electrical Pulse (nsEP) Exposure. Sci Rep 5:15063. doi:

10.1038/srep15063

164. Kotnik T, Kramar P, Pucihar G, et al (2012) Cell membrane electroporation- Part 1:

The phenomenon. IEEE Electr Insul Mag 28:14–23. doi: 10.1109/MEI.2012.6268438

165. Tsong TY (1991) Electroporation of cell membranes. Biophys J 60:297–306. doi:

10.1016/S0006-3495(91)82054-9

166. Weaver JC (1993) Electroporation: a general phenomenon for manipulating cells and

tissues. J Cell Biochem 51:426–435.

167. Kotnik T, Frey W, Sack M, et al (2015) Electroporation-based applications in

biotechnology. Trends Biotechnol. doi: 10.1016/j.tibtech.2015.06.002

168. Mahnič-Kalamiza S, Vorobiev E, Miklavčič D (2014) Electroporation in food

processing and biorefinery. J Membr Biol 247:1279–1304.

169. Toepfl S, Siemer C, Saldaña-Navarro G, Heinz V (2014) Overview of Pulsed Electric

Fields Processing for Food. In: Emerg. Technol. Food Process. Elsevier, pp 93–114

170. Yarmush ML, Golberg A, Serša G, et al (2014) Electroporation-Based Technologies for

Medicine: Principles, Applications, and Challenges. Annu Rev Biomed Eng 16:295–

320. doi: 10.1146/annurev-bioeng-071813-104622

171. Daud AI, DeConti RC, Andrews S, et al (2008) Phase I trial of interleukin-12 plasmid

electroporation in patients with metastatic melanoma. J Clin Oncol 26:5896–5903.

172. Calvet CY, André FM, Mir LM (2014) Dual therapeutic benefit of electroporation-

mediated DNA vaccination in vivo: Enhanced gene transfer and adjuvant activity.

Oncoimmunology 3:e28540. doi: 10.4161/onci.28540

Page 137: Investigating the role of voltage-gated ion channels in

136

173. Trimble CL, Morrow MP, Kraynyak KA, et al (2015) Safety, efficacy, and

immunogenicity of VGX-3100, a therapeutic synthetic DNA vaccine targeting human

papillomavirus 16 and 18 E6 and E7 proteins for cervical intraepithelial neoplasia 2/3:

a randomised, double-blind, placebo-controlled phase 2b trial. Lancet 386:2078–

2088. doi: 10.1016/S0140-6736(15)00239-1

174. Vasan S, Hurley A, Schlesinger SJ, et al (2011) In Vivo Electroporation Enhances the

Immunogenicity of an HIV-1 DNA Vaccine Candidate in Healthy Volunteers. PLoS

One 6:e19252. doi: 10.1371/journal.pone.0019252

175. Serša G, Teissié J, Čemažar M, et al (2015) Electrochemotherapy of tumors as in situ

vaccination boosted by immunogene electrotransfer. Cancer Immunol Immunother

64:1315–1327. doi: 10.1007/s00262-015-1724-2

176. Denet A-R, Vanbever R, Préat V (2004) Skin electroporation for transdermal and

topical delivery. Adv Drug Deliv Rev 56:659–674. doi: 10.1016/j.addr.2003.10.027

177. Zorec B, Préat V, Miklavčič D, Pavšelj N (2013) Active enhancement methods for

intra- and transdermal drug delivery: a review. Zdr Vestn 82:339–56.

178. Al-Sakere B, André F, Bernat C, et al (2007) Tumor Ablation with Irreversible

Electroporation. PLoS One 2:e1135. doi: 10.1371/journal.pone.0001135

179. Gasbarrini A, Campos WK, Campanacci L, Boriani S (2015) Electrochemotherapy to

Metastatic Spinal Melanoma: A Novel Treatment of Spinal Metastasis? Spine (Phila

Pa 1976) 40:E1340–E1346. doi: 10.1097/BRS.0000000000001125

180. Mali B, Jarm T, Snoj M, et al (2013) Antitumor effectiveness of electrochemo-therapy:

A systematic review and meta-analysis. Eur J Surg Oncol 39:4–16.

181. Miklavčič D, Mali B, Kos B, et al (2014) Electrochemotherapy: from the drawing board

into medical practice. Biomed Eng Online 13:29. doi: 10.1186/1475-925X-13-29

182. Bianchi G, Campanacci L, Ronchetti M, Donati D (2016) Electrochemotherapy in the

Page 138: Investigating the role of voltage-gated ion channels in

137

Treatment of Bone Metastases: A Phase II Trial. World J Surg 40:3088–3094. doi:

10.1007/s00268-016-3627-6

183. Hille B (1992) Ionic channels of excitable membranes, 2nd ed. Sinauer Associates,

Sunderland, Mass

184. Garcia PA, Rossmeisl JH, Robertson JL, et al (2012) 7.0-T Magnetic Resonance

Imaging Characterization of Acute Blood-Brain-Barrier Disruption Achieved with

Intracranial Irreversible Electroporation. PLoS One 7:e50482. doi:

10.1371/journal.pone.0050482

185. Garcia PA, Pancotto T, Rossmeisl JH, et al (2011) Non-thermal irreversible

electroporation (N-TIRE) and adjuvant fractionated radiotherapeutic multimodal

therapy for intracranial malignant glioma in a canine patient. Technol Cancer Res

Treat 10:73–83.

186. Hjouj M, Last D, Guez D, et al (2012) MRI Study on Reversible and Irreversible

Electroporation Induced Blood Brain Barrier Disruption. PLoS One 7:e42817. doi:

10.1371/journal.pone.0042817

187. Linnert M, Iversen HK, Gehl J (2012) Multiple brain metastases - current

management and perspectives for treatment with electrochemotherapy. Radiol Oncol.

doi: 10.2478/v10019-012-0042-y

188. Sharabi S, Kos B, Last D, et al (2016) A statistical model describing combined

irreversible electroporation and electroporation-induced blood-brain barrier

disruption. Radiol Oncol 50:28–38.

189. Neal RE, Millar JL, Kavnoudias H, et al (2014) In vivo characterization and numerical

simulation of prostate properties for non-thermal irreversible electroporation

ablation: Characterized and Simulated Prostate IRE. Prostate 74:458–468. doi:

10.1002/pros.22760

Page 139: Investigating the role of voltage-gated ion channels in

138

190. Tschon M, Salamanna F, Ronchetti M, et al (2015) Feasibility of Electroporation in

Bone and in the Surrounding Clinically Relevant Structures: A Preclinical

Investigation. Technol Cancer Res Treat. doi: 10.1177/1533034615604454

191. Lavee J, Onik G, Mikus P, Rubinsky B (2007) A Novel Nonthermal Energy Source for

Surgical Epicardial Atrial Ablation: Irreversible Electroporation. Heart Surg Forum

10:E162–E167. doi: 10.1532/HSF98.20061202

192. Neven K, van Driel V, van Wessel H, et al (2014) Myocardial Lesion Size After

Epicardial Electroporation Catheter Ablation After Subxiphoid Puncture. Circ

Arrhythmia Electrophysiol 7:728–733. doi: 10.1161/CIRCEP.114.001659

193. Xie F, Varghese F, Pakhomov AG, et al (2015) Ablation of Myocardial Tissue With

Nanosecond Pulsed Electric Fields. PLoS One 10:e0144833. doi:

10.1371/journal.pone.0144833

194. Gehl J, Sørensen TH, Nielsen K, et al (1999) In vivo electroporation of skeletal muscle:

threshold, efficacy and relation to electric field distribution. Biochim Biophys Acta -

Gen Subj 1428:233–240. doi: 10.1016/S0304-4165(99)00094-X

195. Satkauskas S, Bureau MF, Puc M, et al (2002) Mechanisms of in Vivo DNA

Electrotransfer: Respective Contributions of Cell Electropermeabilization and DNA

Electrophoresis. Mol Ther 5:133–140. doi: 10.1006/mthe.2002.0526

196. Ayuni EL, Gazdhar A, Giraud MN, et al (2010) In Vivo Electroporation Mediated Gene

Delivery to the Beating Heart. PLoS One 5:e14467. doi: 10.1371/journal.pone.0014467

197. Bulysheva AA, Hargrave B, Burcus N, et al (2016) Vascular endothelial growth factor-

A gene electrotransfer promotes angiogenesis in a porcine model of cardiac ischemia.

Gene Ther 23:649–656. doi: 10.1038/gt.2016.35

198. Li W, Fan Q, Ji Z, et al (2011) The Effects of Irreversible Electroporation (IRE) on

Nerves. PLoS One 6:e18831. doi: 10.1371/journal.pone.0018831

Page 140: Investigating the role of voltage-gated ion channels in

139

199. Onik G, Mikus P, Rubinsky B (2007) Irreversible electroporation: implications for

prostate ablation. Technol Cancer Res Treat 6:295–300.

200. Casciola M, Xiao S, Pakhomov AG (2017) Damage-free peripheral nerve stimulation

by 12-ns pulsed electric field. Sci Rep 7:10453. doi: 10.1038/s41598-017-10282-5

201. Schoellnast H, Monette S, Ezell PC, et al (2011) Acute and subacute effects of

irreversible electroporation on nerves: experimental study in a pig model. Radiology

260:421–427. doi: 10.1148/radiol.11103505

202. Nevian T, Helmchen F (2007) Calcium indicator loading of neurons using single-cell

electroporation. Pflügers Arch - Eur J Physiol 454:675–688. doi: 10.1007/s00424-

007-0234-2

203. van Driel VJHM, Neven K, van Wessel H, et al (2015) Low vulnerability of the right

phrenic nerve to electroporation ablation. Hear Rhythm 12:1838–1844. doi:

10.1016/j.hrthm.2015.05.012

204. Arena CB, Davalos R V. (2012) Advances in Therapeutic Electroporation to Mitigate

Muscle Contractions. J Membr Sci Technol. doi: 10.4172/2155-9589.1000e102

205. Golberg A, Rubinsky B (2012) Towards Electroporation Based Treatment Planning

Considering Electric Field Induced Muscle Contractions. Technol Cancer Res Treat

11:189–201. doi: 10.7785/tcrt.2012.500249

206. Miklavčič D, Pucihar G, Pavlovec M, et al (2005) The effect of high frequency electric

pulses on muscle contractions and antitumor efficiency in vivo for a potential use in

clinical electrochemotherapy. Bioelectrochemistry 65:121–128. doi:

10.1016/j.bioelechem.2004.07.004

207. Županič A, Ribarič S, Miklavčič D (2007) Increasing the repetition frequency of

electric pulse delivery reduces unpleasant sensations that occur in

electrochemotherapy. Neoplasma 54:246–250.

Page 141: Investigating the role of voltage-gated ion channels in

140

208. Ball C, Thomson KR, Kavnoudias H (2010) Irreversible Electroporation: A New

Challenge in “Out of Operating Theater” Anesthesia. Anesth Analg 110:1305–1309.

doi: 10.1213/ANE.0b013e3181d27b30

209. Deodhar A, Dickfeld T, Single GW, et al (2011) Irreversible electroporation near the

heart: ventricular arrhythmias can be prevented with ECG synchronization. AJR Am J

Roentgenol 196:W330-335. doi: 10.2214/AJR.10.4490

210. Mali B, Gorjup V, Edhemovic I, et al (2015) Electrochemotherapy of colorectal liver

metastases - an observational study of its effects on the electrocardiogram. Biomed

Eng Online 14:S5. doi: 10.1186/1475-925X-14-S3-S5

211. Mali B, Jarm T, Corovic S, et al (2008) The effect of electroporation pulses on

functioning of the heart. Med Biol Eng Comput 46:745–757. doi: 10.1007/s11517-008-

0346-7

212. Jiang N, Cooper BY (2011) Frequency-dependent interaction of ultrashort E-fields

with nociceptor membranes and proteins. Bioelectromagnetics 32:148–163. doi:

10.1002/bem.20620

213. Nene D, Jiang N, Rau KK, et al (2006) Nociceptor activation and damage by pulsed E-

fields. doi: 10.1117/12.665181

214. Spitzner M, Ousingsawat J, Scheidt K, et al (2006) Voltage-gated K+ channels support

proliferation of colonic carcinoma cells. FASEB J 21:35–44. doi: 10.1096/fj.06-

6200com

215. Wolff C, Fuks B, Chatelain P (2003) Comparative study of membrane potential-

sensitive fluorescent probes and their use in ion channel screening assays. J Biomol

Screen 8:533–543. doi: 10.1177/1087057103257806

216. Lalik PH, Krafte DS, Volberg WA, Ciccarelli RB (1993) Characterization of

endogenous sodium channel gene expressed in Chinese hamster ovary cells. Am J

Page 142: Investigating the role of voltage-gated ion channels in

141

Physiol Cell Physiol 264:C803-9.

217. Skryma R, Prevarskaya N, Vacher P, Dufy B (1994) Voltage-dependent Ca2+ channels

in Chinese hamster ovary (CHO) cells. FEBS 349:289–294.

218. Gamper N, Stockand JD, Shapiro MS (2005) The use of Chinese hamster ovary (CHO)

cells in the study of ion channels. J Pharmacol Toxicol Methods 51:177–185. doi:

10.1016/j.vascn.2004.08.008

219. Kandel ER, Schwartz JH, Jessell TM, et al (2012) Principles of neural science, 5th ed.

McGraw-Hill, Education

220. Frandsen SK, Kruger MB, Mangalanathan UM, et al (2017) Normal and Malignant

Cells Exhibit Differential Responses to Calcium Electroporation. Cancer Res 77:1–14.

doi: 10.1158/0008-5472.CAN-16-1611

221. Frandsen SK, McNeil AK, Novak I, et al (2016) Difference in Membrane Repair

Capacity Between Cancer Cell Lines and a Normal Cell Line. J Membr Biol 249:569–

576. doi: 10.1007/s00232-016-9910-5

222. Fairless R, Beck A, Kravchenko M, et al (2013) Membrane Potential Measurements of

Isolated Neurons Using a Voltage-Sensitive Dye. PLoS One 8:e58260. doi:

10.1371/journal.pone.0058260

223. Rols MP, Delteil C, Golzio M, Teissié J (1998) Control by ATP and ADP of voltage-

induced mammalian-cell-membrane permeabilization, gene transfer and resulting

expression. Eur J Biochem 254:382–388.

224. Chen W, Zhongsheng Z, Lee RC (2006) Supramembrane potential-induced

electroconformational changes in sodium channel proteins: A potential mechanism

involved in electric injury. Burns 32:52–59. doi: 10.1016/j.burns.2005.08.008

225. Levin M, Stevenson CG (2012) Regulation of Cell Behavior and Tissue Patterning by

Bioelectrical Signals: Challenges and Opportunities for Biomedical Engineering. Annu

Page 143: Investigating the role of voltage-gated ion channels in

142

Rev Biomed Eng 14:295–323. doi: 10.1146/annurev-bioeng-071811-150114

226. Dermol J, Miklavčič D (2014) Predicting electroporation of cells in an inhomogeneous

electric field based on mathematical modeling and experimental CHO-cell

permeabilization to propidium iodide determination. Bioelectrochemistry 100:52–61.

doi: 10.1016/j.bioelechem.2014.03.011

227. Pucihar G, Kotnik T, Teissié J, Miklavčič D (2007) Electropermeabilization of dense

cell suspensions. Eur Biophys J 36:172–185.

228. Susil R, Šemrov D, Miklavčič D (1998) Electric field induced transmembrane potential

depends on cell density and organization. Electro- and Magnetobiology 17:391–399.

229. Čemažar M, Jarm T, Miklavčič D, et al (1998) Effect of electric-field intensity on

electropermeabilization and electrosensitivity of various tumor-cell lines in vitro.

Electro- and Magnetobiology 17:263–272.

230. Moen EK, Ibey BL, Beier HT, Armani AM (2016) Quantifying pulsed electric field-

induced membrane nanoporation in single cells. Biochim Biophys Acta - Biomembr

1858:2795–2803. doi: 10.1016/j.bbamem.2016.08.007

231. Muller KJ, Sukhorukov VL, Zimmermann U (2001) Membrane Biology Reversible

Electropermeabilization of Mammalian Cells by High-Intensity , Ultra-Short. J

Membr Biol 184:161–170. doi: 10.1007/s00232-001-0084-3

232. Satkauskas S, Ruzgys P, Venslauskas MS (2012) Towards the mechanisms for efficient

gene transfer into cells and tissues by means of cell electroporation. Expert Opin Biol

Ther 12:275–286. doi: 10.1517/14712598.2012.654775

233. Kinosita KJ, Tsong TY (1977) Voltage-induced pore formation and hemolysis of

human erythrocytes. Biochim Biophys Acta 471:227–242.

234. Sale AJH, Hamilton WA (1968) Effects of high electric fields on micro-organisms: III.

Lysis of erythrocytes and protoplasts. Biochim Biophys Acta - Biomembr 163:37–43.

Page 144: Investigating the role of voltage-gated ion channels in

143

doi: 10.1016/0005-2736(68)90030-8

235. Yang M, Brackenbury WJ (2013) Membrane potential and cancer progression. Front

Physiol 4:1–10. doi: 10.3389/fphys.2013.00185

236. Voets T, Owsianik G, Janssens A, et al (2007) TRPM8 voltage sensor mutants reveal a

mechanism for integrating thermal and chemical stimuli. Nat Chem Biol 3:174–182.

doi: 10.1038/nchembio862

237. Fernández JA, Skryma R, Bidaux G, et al (2011) Voltage- and cold-dependent gating of

single TRPM8 ion channels. J Gen Physiol 137:173–195. doi: 10.1085/jgp.201010498

238. Raddatz N, Castillo JP, Gonzalez C, et al (2014) Temperature and Voltage Coupling to

Channel Opening in Transient Receptor Potential Melastatin 8 ( TRPM8 ) *. J Biol

Chem 289:35438–35454. doi: 10.1074/jbc.M114.612713

239. Belehradek M, Domenge C, Luboinski B, et al (1993) Electrochemotherapy , a New

Antitumor Treatment First Clinical Phase I-II Trial. Cancer 72:3694–3700.

240. Falk H, Matthiessen LW, Wooler G, Gehl J (2017) Calcium electroporation for

treatment of cutaneous metastases ; a randomized double- blinded phase II study ,

comparing the effect of calcium electroporation with electrochemotherapy. Acta Oncol

(Madr) 1–9. doi: 10.1080/0284186X.2017.1355109

241. Gehl J, Geertsen PF (2000) Efficient palliation of haemorrhaging malignant

melanoma skin metastases by electrochemotherapy. Melanoma Res 10:585–589.

242. Boinagrov D, Loudin J, Palanker D (2010) Strength-Duration Relationship for

Extracellular Neural Stimulation: Numerical and Analytical Models. J Neurophysiol

104:2236–2248. doi: 10.1152/jn.00343.2010

243. Brunel N, Van Rossum MCW (2007) Quantitative investigations of electrical nerve

excitation treated as polarization. Biol Cybern 97:341–349. doi: 10.1007/s00422-007-

0189-6

Page 145: Investigating the role of voltage-gated ion channels in

144

244. Chen W, Han Y, Chen Y, Astumian D (1998) Electric field-induced functional

reductions in the K+ channels mainly resulted from supramembrane potential-

mediated electroconformational changes. Biophys J 75:196–206. doi: 10.1016/S0006-

3495(98)77506-X

245. Chen W (2004) Supra-physiological membrane potential induced conformational

changes in K+ channel conducting system of skeletal muscle fibers.

Bioelectrochemistry 62:47–56. doi: 10.1016/j.bioelechem.2003.10.006

Page 146: Investigating the role of voltage-gated ion channels in

145

List of publications during Ph.D. candidature

Journal articles

1. Burke RC, Bardet SM, Carr L, et al (2017) Nanosecond pulsed electric fields

depolarize transmembrane potential via voltage-gated K+, Ca2+ and TRPM8 channels

in U87 glioblastoma cells. BBA - Biomembr 1859:2040–2050. doi:

10.1016/j.bbamem.2017.07.004

2. Carr L, Bardet SM, Burke RC, et al (2017) Calcium-independent disruption of

microtubule dynamics by nanosecond pulsed electric fields in U87 human

glioblastoma cells. Sci Rep. doi: 10.1038/srep41267

3. Moreau D, Lefort C, Burke R, et al (2015) Rhodamine B as an optical thermometer in

cells focally exposed to infrared laser light or nanosecond pulsed electric fields.

Biomed Opt Express 6:713–718. doi: 10.1364/BOE.6.004105

4. Dermol, Janja, Miklavcic, Damijan, Rebersek, Matej, et al (2017) Plasma membrane

depolarization and permeabilization due to electric pulses in cell lines of different

excitability. Bioelectrochemistry (in review)

Page 147: Investigating the role of voltage-gated ion channels in

146

Presentations at International Conferences

1. Burke R., Romanenko S., Moreau D., Arnaud-Cormos D., Leveque P., O'Connor R.P.

"Application of a voltage sensitive dye to study the effect of nanosecond pulsed

electric fields (nsPEF) on membrane potential in human U87 glioblastoma cells." In

BioEM 2015, Joint Meeting of the BioElectroMagnetics Society and the European

BioElectromagnetics Association, Pacific Grove, CA, USA, 14 - 19 June 2015

2. Burke, R., Moreau, David., Arnaud-Cormos, Delia., Leveque, Philippe., O’Connor,

Rodney. “Voltage-gated ion channel antagonists inhibit nsPEF-induced membrane

depolarization in U87 glioblastoma cells.” In EBTT 2015 proceedings, Ljubljana,

Slovenia, November 15 – 21, 2015.

3. Moreau, David., Lefort, Claire., Burke, Ryan., Leveque, Philippe., O’Connor,

Rodney. “Thermal effect of nanosecond pulsed electric fields.” In EBTT 2015

proceedings, Ljubljana, Slovenia, November 15 – 21, 2015.

4. Moreau, David., Lefort, Claire., Burke, Ryan., Leveque, Philippe., O’Connor,

Rodney. “Thermal imaging with Rhodamine B in cells exposed to electromagnetic

radiation.” In BioEM 2016, joint meeting of the Bioelectromagnetics Society and the

European Bioelectromagnetics Association, Ghent, Belgium, June 5 – 10, 2016.

5. Dermol, Janja., Arnaud-Cormos, Delia., Bardet, Sylvia., Burke, Ryan., Leveque,

Philippe., Mekuc, Primoz., Miklavcic, Damijan. “Cell membrane depolarization and

permeability of three cell lines of different excitability.” In BES 2017, XXIV

International symposium on Bioelectrochemistry and Bioenergetics of the

Bioelectrochemical Society, Lyon, France, July 3 – 7, 2017.

Page 148: Investigating the role of voltage-gated ion channels in

147

6. Frandsen, Stine., Burke, Ryan., Arnaud-Cormos, Delia., Bardet, Sylvia., Leveque,

Philippe. “Effect of in vitro calcium-electroporation on mitochondrial membrane

potential.” WC2017, 2nd World Congress on Electroporation and pulsed electric fields

in biology, medicine and food & environmental technologies, Norfolk, Virginia, USA,

September 24 – 28, 2017.